0711.2703/11.tex
1: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
2: \documentclass[11pt]{sat}
3: \usepackage{amsmath,amsthm,amscd,amssymb}
4: \usepackage{latexsym}
5: 
6: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7: %%%%%%%%%%%%%%%%%%%% INSERT YOUR MACROS HERE %%%%%%%%%%%%%%%%%%%%%%%%%
8: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
9: 
10: %%%%%%%%%%%%% fonts/sets %%%%%%%%%%%%%%%%%%%%%%%
11: \newcommand{\R}{{\mathbb{R}}}
12: \newcommand{\Z}{{\mathbb{Z}}}
13: \newcommand{\C}{{\mathbb{C}}}
14: \newcommand{\D}{{\mathbb{D}}}
15: \newcommand{\h}{{\mathcal{H}}}
16: \newcommand{\pol}{{\mathcal{P}}}
17: \newcommand{\calH}{{\mathcal H}}
18: \newcommand{\calL}{{\mathcal L}}
19: \newcommand{\calM}{{\mathcal M}}
20: \newcommand{\calV}{{\mathcal V}}
21: 
22: 
23: \newcommand{\bbC}{{\mathbb{C}}}
24: \newcommand{\bbD}{{\mathbb{D}}}
25: \newcommand{\bbL}{{\mathbb{L}}}
26: \newcommand{\bbR}{{\mathbb{R}}}
27: \newcommand{\bbS}{{\mathbb{S}}}
28: \newcommand{\bbT}{{\mathbb{T}}}
29: \newcommand{\bbU}{{\mathbb{U}}}
30: \newcommand{\bbZ}{{\mathbb{Z}}}
31: 
32: \allowdisplaybreaks
33: \numberwithin{equation}{section}
34: 
35: \newtheorem{theorem}{Theorem}[section]
36: \newtheorem{proposition}[theorem]{Proposition}
37: \newtheorem{lemma}[theorem]{Lemma}
38: \newtheorem{corollary}[theorem]{Corollary}
39: \theoremstyle{definition}
40: %\newtheorem{definition}[theorem]{Definition}
41: \newtheorem{example}[theorem]{Example}
42: \newtheorem{conjecture}[theorem]{Conjecture}
43: \newtheorem{xca}[theorem]{Exercise}
44: 
45: \theoremstyle{remark}
46: \newtheorem*{remark}{Remark}
47: \newtheorem*{remarks}{Remarks}
48: \newtheorem*{definition}{Definition}
49: 
50: % Absolute value notation
51: \newcommand{\abs}[1]{\lvert#1\rvert}
52: 
53: 
54: \newcommand{\norm}[1]{\lVert#1\rVert}
55: \newcommand{\snorm}[1]{{ \pmb\lvert}#1{ \pmb\rvert}}
56: 
57: \DeclareMathOperator{\Tr}{Tr}
58: \DeclareMathOperator{\diam}{diam}
59: \DeclareMathOperator{\Ima}{Im}
60: \DeclareMathOperator{\Span}{span}
61: \DeclareMathOperator{\Ran}{Ran}
62: \DeclareMathOperator{\lc}{linear\  combination\  of}
63: 
64: %\newcommand{\ang}[1]{\langle \langle#1\rangle \rangle}
65: \newcommand{\ang}[1]{\langle\! \langle#1\rangle\! \rangle}
66: \newcommand{\vect}[2]{\begin{pmatrix}#1 \\ #2 \end{pmatrix}}
67: 
68: \def\jap#1{\langle #1 \rangle}
69: 
70: \newcommand{\lb}{\label}
71: \newcommand{\bdone}{{\boldsymbol{1}}}
72: \newcommand{\bdzero}{{\boldsymbol{0}}}
73: \newcommand{\bddot}{{\boldsymbol{\cdot}}}
74: \newcommand{\singc}{\text{\rm{sc}}}
75: \newcommand{\supp}{\text{\rm{supp}}}
76: \newcommand{\ess}{\text{\rm{ess}}}
77: \newcommand{\s}{\text{\rm{s}}}
78: \newcommand{\ol}{\overline}
79: \newcommand{\f}{\frac}
80: \newcommand{\dott}{\,\cdot\,}
81: \newcommand{\cvh}{\text{\rm{cvh}}}
82: \newcommand{\tr}{\text{\rm{tr}}}
83: \newcommand{\dist}{\text{\rm{dist}}}
84: \newcommand{\ti}{\tilde  }
85: \newcommand{\wti}{\widetilde  }
86: 
87: \let\det=\undefined\DeclareMathOperator{\det}{det}
88: 
89: %%%%%%%%%%%%% marginal warnings %%%%%%%%%%%%%%%%
90: % ON:
91: \newcommand{\TK}{{\marginpar{x-ref?}}}
92: % OFF:
93: %\newcommand{\TK}{}
94: 
95: %  Rowan's unspaced list
96: %
97: \newcounter{smalllist}
98: \newenvironment{SL}{\begin{list}{{\rm\roman{smalllist})}}{%
99: \setlength{\topsep}{0mm}\setlength{\parsep}{0mm}\setlength{\itemsep}
100: {0mm}%
101: \setlength{\labelwidth}{2em}\setlength{\leftmargin}{2em}\usecounter
102: {smalllist}%
103: }}{\end{list}}
104: 
105: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
106: %%%%%%%%%%%%%%%%%%%%%% END OF YOUR MACROS  %%%%%%%%%%%%%%%%%%%%%%%%%%%
107: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
108: 
109: \sloppy
110: 
111: %\begin{document}
112: 
113: \title{The Analytic Theory of Matrix
114: Orthogonal Polynomials}
115: \def\shorttitle{Matrix Orthogonal Polynomials}
116: 
117: \author{David Damanik, Alexander Pushnitski, and Barry Simon}
118: \def\shortauthor{D.\ Damanik, A.\ Pushnitski, B.\ Simon }
119: 
120: 
121: \def\versiondate{January 30, 2008}
122: 
123: \def\abstracttext{We survey the analytic theory of matrix orthogonal polynomials.}
124: 
125: \def\MSCnumbers{42C05, 47B36, 30C10}
126: 
127: \def\keywords{orthogonal polynomials, matrix-valued measures, block
128: Jacobi matrices, block CMV matrices}
129: 
130: 
131: 
132: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
133: %%%%%%%%%%%%%%%%%%%%%%% FOR EDITORS %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
134: %%%%%%%%%%%%%%%%%% DO NOT MODIFY THESE %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
135: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
136: 
137: %%%%%%%%%%%% Initializing (to be done by the editors)
138: \def\startpagenumber{1}
139: \def\volumenumber{4}
140: \def\year{2008}
141: 
142: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
143: 
144: \setcounter{page}{\startpagenumber}
145: \pagestyle{myheadings}
146: \newcommand{\beginddoc}{\begin{document}
147: \maketitle
148: \begin{abstract}
149: \abstracttext
150: \vskip1pt MSC: \MSCnumbers
151: \ifx\keywords\empty\else\vskip1pt keywords: \keywords\fi
152: \end{abstract}
153: \insert\footins{\scriptsize
154: \medskip
155: \baselineskip 8pt
156: \leftline{Surveys in Approximation Theory}
157: \leftline{Volume \volumenumber, \year.
158: pp.~\thepage--\pageref{endpage}.}
159: \leftline{\copyright\ \year\ Surveys in Approximation Theory.}
160: \leftline{ISSN 1555-578X}
161: \leftline{All rights of reproduction in any form reserved.}
162: \smallskip
163: \par\allowbreak}
164: \tableofcontents}
165: \renewcommand\rightmark{\ifodd\thepage{\it \hfill\shorttitle\hfill}\else {\it \hfill\shortauthor\hfill}\fi}
166: \markboth{{\it \shortauthor}}{{\it \shorttitle}}
167: \markright{{\it \shorttitle}}
168: \def\endddoc{\label{endpage}\end{document}}
169: \date{{\small \versiondate}}
170: \setlength\oddsidemargin{0pc}
171: \setlength\evensidemargin{0pc}
172: \setlength\topmargin{0in}
173: \setlength\textwidth{6.5in}
174: \setlength\textheight{8.6in}
175: \beginddoc
176: 
177: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
178: %%%%%%%%%%%%%%%%%%%%%% YOUR MATHEMATICS  %%%%%%%%%%%%%%%%%%%%%%%%%%%%%
179: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
180: 
181: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
182: %%%%%%%%%%%%%%% INSERT YOUR ARTICLE HERE USING %%%%%%%%%%%%%%%%%%%%%%%
183: %%%%%%%%%%%%%%%%%% STANDARD LATEX COMMANDS %%%%%%%%%%%%%%%%%%%%%%%%%%%
184: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
185: 
186: %\maketitle
187: 
188: %\tableofcontents
189: 
190: \section{Introduction}
191: 
192: \subsection{Introduction and Overview} \lb{s1.1}
193: 
194: Orthogonal polynomials on the real line (OPRL) were developed in
195: the nineteenth century and orthogonal polynomials on the unit
196: circle (OPUC) were initially developed around 1920 by Szeg\H{o}.
197: Their matrix analogues are of much more recent vintage. They were
198: originally developed in the MOPUC case indirectly in the study of
199: prediction theory \cite{HL1,HL2, Kol,Krein1,Krein2,Lev,Wie} in the
200: period 1940--1960. The connection to OPUC in the scalar case was
201: discovered by Krein \cite{Krein1}. Much of the theory since is in
202: the electrical engineering literature \cite
203: {DG92,DGK,DGK3,DGK79,DGK2,DGK81,Kai74,Kai87,K6.1,KVM,YK}; see also
204: \cite{Ger81,Ge92,GC,GW07,MarRod}.
205: 
206: The corresponding real line theory (MOPRL) is still more recent:
207: Following early work of Krein \cite{Krein3} and Berezan'ski
208: \cite{Ber} on block Jacobi matrices, mainly as applied to
209: self-adjoint extensions, there was a seminal paper of
210: Aptekarev--Nikishin \cite{AN} and a flurry of papers since the
211: 1990s \cite{Berg,Berg-Duran1,Berg-Duran2,CFMV,CMV05,CMV07,CG05,CG,CG07,CH,CGR,DGIM,DS,Dur1,Dur95,
212: Dur96,Dur6,Dur99,Dur7,Dur8,DuDa,DD,DDe,DG04,DG05a,DG05b,DG05c,DG05d,DG06,DG07a,DG07b,DdI07a,
213: DdI07b,DL,DL97a,DL97b, DL00,DL04,DL07,DLS,DP02,DP,DVA,Dyu,Fu,Ger82,Gru,Gru07,
214: GI07a,GI07b,GI03,GPT01,GPT02a,GPT02,GPT03,GPT04,GPT05,GT07,LR,Lop,Lop01,
215: MS,MY,MZ,M05a,M05b,M05c,PT04,PT07a,PT07b,PR,RT,Ros,SVI,Tirao,
216: Yak,YM02,YMP01,YMP02,Zyg}; see also \cite{BB83}.
217: 
218: There is very little on the subject in monographs
219: --- the more classical ones (e.g.,
220: \cite{Chi,FrB,Ger77,Szb}) predate most of the subject; see,
221: however, Atkinson \cite[Section~6.6]{Atk}. Ismail \cite {IsmBk} has no
222: discussion and Simon \cite{S,S2} has a single section! Because of
223: the use of MOPRL in \cite{DKSppt}, we became interested in the
224: subject and, in particular, we needed some basic results for that
225: paper which we couldn't find in the literature or which, at least,
226: weren't very accessible. Thus, we decided to produce this
227: comprehensive review that we hope others will find useful.
228: 
229: As with the scalar case, the subject breaks into two parts,
230: conveniently called the analytic theory (general structure
231: results) and the algebraic theory (the set of non-trivial
232: examples). This survey deals entirely with the analytic theory. We
233: note, however, that one of the striking developments in recent
234: years has been the discovery that there are rich classes of
235: genuinely new MOPRL, even at the classical level of Bochner's
236: theorem; see \cite{CG,DG04,DuLo4,DuLo5,Gru,GPT02,GPT03,GPT04,GPT05,GT07,PT07a,RT}
237: and the forthcoming monograph \cite{DGPT} for further
238: discussion of this algebraic side.
239: 
240: In this introduction, we will focus mainly on the MOPRL case. For
241: scalar OPRL, a key issue
242: is the passage from measure to monic OPRL, then to normalized OPRL,
243: and finally to
244: Jacobi parameters. There are no choices in going from measure to
245: monic OP, $P_n(x)$.
246: They are determined by
247: \begin{equation} \lb{1.1}
248: P_n(x) = x^n + \text{lower order}, \qquad \jap{x^j, P_n}=0 \quad j=1,
249: \dots, n-1.
250: \end{equation}
251: 
252: However, the basic condition on the orthonormal polynomials, namely,
253: \begin{equation} \lb{1.2}
254: \jap{p_n,p_m} =\delta_{nm}
255: \end{equation}
256: does not uniquely determine the $p_n(x)$. The standard choice is
257: $$
258: p_n(x) =  \f{P_n(x)}{\norm{P_n}}.
259: $$
260: However, if
261: $\theta_0, \theta_1, \dots$ are arbitrary real numbers,
262: then
263: \begin{equation} \lb{1.3}
264: \ti p_n(x) =  \f{e^{i\theta_n} P_n(x)}{\norm{P_n}}
265: \end{equation}
266: also obey \eqref{1.2}. If the recursion coefficients (aka Jacobi
267: parameters), are defined via
268: \begin{equation} \lb{1.4}
269: xp_n =a_{n+1} p_{n+1} + b_{n+1} p_n + a_n p_{n-1},
270: \end{equation}
271: then the choice \eqref{1.3} leads to
272: \begin{equation} \lb{1.5}
273: \ti b_n =b_n, \qquad \ti a_n =e^{i\theta_n} a_n e^{-i\theta_{n-1}}.
274: \end{equation}
275: 
276: The standard choice is, of course, most natural here; for example, if
277: \begin{equation} \lb{1.6}
278: p_n(x) = \kappa_n x^n + \text{lower order},
279: \end{equation}
280: then $a_n>0$ implies $\kappa_n >0$. It would be crazy to make any
281: other choice.
282: 
283: For MOPRL, these choices are less clear. As we will explain in
284: Section~\ref{s1.2}, there are
285: now two matrix-valued ``inner products" formally written as
286: \begin{align}
287: \ang{f,g}_R &= \int f(x)^\dagger\, d\mu(x) g(x), \lb{1.7} \\
288: \ang{f,g}_L &= \int g(x)\, d\mu(x) f (x)^\dagger, \lb{1.8}
289: \end{align}
290: where now $\mu$ is a matrix-valued measure and  ${}^\dagger$
291: denotes the adjoint, and corresponding two sets of monic OPRL:
292: $P_n^R(x)$ and $P_n^L(x)$. The orthonormal polynomials are
293: required to obey
294: \begin{equation} \lb{1.9}
295: \ang{p_n^R,p_m^R}_R = \delta_{nm}\bdone.
296: \end{equation}
297: The analogue of \eqref{1.3} is
298: \begin{equation} \lb{1.10z}
299: \ti p_n^R(x) = P_n^R(x) \ang{P_n^R,P_n^R}^{-1/2} \sigma_n
300: \end{equation}
301: for a unitary $\sigma_n$. For the immediately following, use $p_n^R$
302: to be
303: the choice $\sigma_n\equiv 1$. For any such choice, we have a
304: recursion relation,
305: \begin{equation} \lb{1.10x}
306: xp_n^R(x) =p_{n+1}^R(x) A_{n+1}^\dagger + p_n^R(x) B_{n+1} + p_{n-1}^R
307: (x) A_n
308: \end{equation}
309: with the analogue of \eqref{1.5} (comparing $\sigma_n\equiv
310: \bdone$ to general $\sigma_n$)
311: \begin{equation} \lb{1.10}
312: \ti B_n =\sigma_n^\dagger B_n \sigma_n \qquad \ti A_n =\sigma_{n-1}^
313: \dagger A_n\sigma_n.
314: \end{equation}
315: 
316: 
317: The obvious analogue of the scalar case is to pick $\sigma_n\equiv
318: \bdone$, which makes $\kappa_n$ in
319: \begin{equation} \lb{1.11x}
320: p_n^R(x) = \kappa_n x^n + \text{lower order}
321: \end{equation}
322: obey $\kappa_n >0$. Note that \eqref{1.10x} implies
323: \begin{equation} \lb{1.11}
324: \kappa_n = \kappa_{n+1} A_{n+1}^\dagger
325: \end{equation}
326: or, inductively,
327: \begin{equation} \lb{1.13}
328: \kappa_n = (A_n^\dagger \dots A_1^\dagger)^{-1}.
329: \end{equation}
330: In general, this choice does not lead to $A_n$ positive or even
331: Hermitian. Alternatively, one can pick $\sigma_n$ so $\ti A_n$ is
332: positive. Besides these two ``obvious" choices, $\kappa_n > 0$ or
333: $A_n >0$, there is a third that $A_n$ be lower triangular that, as
334: we will see in Section~\ref{s1.4}, is natural. Thus, in the study
335: of MOPRL one needs to talk about equivalent sets of $p_n^R$ and of
336: Jacobi parameters, and this is a major theme of Chapter~\ref{s2}.
337: Interestingly enough for MOPUC, the commonly picked choice
338: equivalent to $A_n>0$ (namely, $\rho_n >0$) seems to suffice for
339: applications. So we do not discuss equivalence classes for MOPUC.
340: 
341: Associated to a set of matrix Jacobi parameters is a block Jacobi
342: matrix, that is, a matrix which
343: when written in $l\times l$ blocks is tridiagonal; see \eqref{b6.1}
344: below.
345: 
346: In Chapter~\ref{s2}, we discuss the basics of MOPRL while Chapter~\ref
347: {s3} discusses MOPUC.
348: Chapter~\ref{s4} discusses the Szeg\H{o} mapping connection of MOPUC
349: and MOPRL. Finally,
350: Chapter~\ref{s5} discusses the extension of the theory of regular OPs
351: \cite{StT} to MOPRL.
352: 
353: While this is mainly a survey, it does have numerous new results, of
354: which we mention:
355: \begin{SL}
356: \item[(a)] The clarification of equivalent Jacobi parameters and
357: several new theorems (Theorems~\ref{l.b4} and \ref{l.b5}).
358: \item[(b)] A new result (Theorem~\ref{T2.18H}) on the order of
359: poles or zeros of $m(z)$ in terms of eigenvalues of $J$ and the
360: once stripped $J^{(1)}$. \item[(c)] Formulas for the resolvent in
361: the MOPRL (Theorem~\ref {T2.27}) and MOPUC (Theorem~\ref{T3.21})
362: cases. \item[(d)] A theorem on zeros of $\det(\Phi_n^R)$
363: (Theorem~\ref{T3.6A}) and eigenvalues of a cutoff CMV matrix
364: (Theorem~\ref{T3.6D}). \item[(e)] A new proof of the Geronimus
365: relations (Theorem~\ref{T4.2}). \item[(f)] Discussion of regular
366: MOPRL (Chapter~\ref{s5}).
367: \end{SL}
368: 
369: \smallskip
370: There are numerous open questions and conjectures in this paper, of
371: which we mention:
372: \begin{SL}
373: \item[(1)] We prove that type~1 and type~3 Jacobi parameters in
374: the Nevai class have $A_n\to\bdone$ but do not know if this is
375: true for type~2 and, if so, how to prove it.
376: 
377: \item[(2)] Determine which monic matrix polynomials, $\Phi$, can
378: occur as monic MOPUC.
379: We know $\det(\Phi(z))$ must have all of its zeros in the unit
380: disk in $\C$, but unlike the scalar
381: case where this is sufficient, we do not know necessary and
382: sufficient conditions.
383: 
384: \item[(3)] Generalize Khrushchev theory
385: \cite{Kh2000,Khr,KhGo} to MOPUC; see Section~\ref{s3.13}.
386: 
387: \item[(4)] Provide a proof of Geronimus relations for MOPUC that
388: uses the theory of canonical moments \cite{DS}; see the discussion
389: at the start of Chapter~\ref{s4}.
390: 
391: \item[(5)] Prove Conjecture~\ref{Con5.7} extending a result of
392: Stahl--Totik \cite{StT} from OPRL to MOPRL.
393: \end{SL}
394: 
395: 
396: \subsection{Matrix-Valued Measures} \lb{s1.2}
397: 
398: Let $\calM_l$ denote the ring of all $l\times l$ complex-valued
399: matrices; we denote by
400: $\alpha^\dagger$ the Hermitian conjugate of $\alpha\in \calM_l$.
401: (Because of the use of
402: $\,^*\,$ for Szeg\H{o} dual in the theory of OPUC, we do not use it
403: for adjoint.)
404: For $\alpha\in\calM_l$, we denote by $\norm{\alpha}$ its Euclidean
405: norm  (i.e., the norm of $\alpha$ as a linear operator on $\C^l$ with the
406: usual Euclidean norm).
407: Consider the
408: set $\pol$ of all polynomials in $z\in\C$ with coefficients from $
409: \calM_l$. The set $\pol$
410: can be considered either as a right or as a left module over $\calM_l
411: $; clearly, conjugation
412: makes the left and right structures isomorphic. For $n=0,1,\dots$, $
413: \pol_n$ will denote those
414: polynomials in $\pol$ of degree at most $n$. The set $\mathcal{V}$
415: denotes the set of all
416: polynomials in $z \in \C$ with coefficients from $\C^l$. The standard
417: inner product in
418: $\C^l$ is denoted by $\langle\cdot, \cdot\rangle_{\C^l}$.
419: 
420: A matrix-valued measure, $\mu$, on $\R$ (or $\C$) is the
421: assignment of a positive semi-definite $l\times l$ matrix $\mu(X)$
422: to every Borel set $X$ which is countably additive. We will
423: usually normalize it by requiring
424: \begin{equation} \lb{1.20}
425: \mu(\R)=\bdone
426: \end{equation}
427: (or $\mu(\C)=\bdone$) where $\bdone$ is the $l\times l$ identity
428: matrix. (We use $\bdone$ in general for an identity operator,
429: whether in $\calM_l$ or in the operators on some other Hilbert
430: space, and $\bdzero$ for the zero operator or matrix.) Normally,
431: our measures for MOPRL will have compact support and, of course,
432: our measures for MOPUC will be supported on all or part of
433: $\partial\D$ ($\D$ is the unit disk in $\C$).
434: 
435: Associated to any such measures is a scalar measure
436: \begin{equation} \lb{1.21}
437: \mu_\tr(X) =\Tr(\mu(X))
438: \end{equation}
439: (the trace normalized by $\Tr(\bdone)=l$). $\mu_\tr$ is normalized
440: by $\mu_\tr(\R)=l$.
441: 
442: Applying the Radon--Nikodym theorem to the matrix elements of $\mu$,
443: we see there is a
444: positive semi-definite matrix function $M_{ij}(x)$ so
445: \begin{equation} \lb{1.22}
446: d\mu_{ij}(x)=M_{ij}(x)\, d\mu_\tr (x).
447: \end{equation}
448: Clearly, by \eqref{1.21},
449: \begin{equation} \lb{1.23}
450: \Tr(M(x)) =1
451: \end{equation}
452: for $d\mu_\tr$-a.e.\ $x$. Conversely, any scalar measure with
453: $\mu_\tr (\R)=l$ and positive semi-definite matrix-valued function
454: $M$ obeying \eqref{1.23} define a matrix-valued measure normalized
455: by \eqref{1.20}.
456: 
457: Given $l\times l$ matrix-valued functions $f,g$, we define the $l
458: \times l$ matrix $\ang{f,g}_R$ by
459: \begin{equation} \lb{1.24}
460: \ang{f,g}_R=\int f(x)^\dagger M(x)g(x)\, d\mu_\tr (x),
461: \end{equation}
462: that is, its $(j,k)$ entry is
463: \begin{equation} \lb{1.25}
464: \sum_{nm} \int \ol{f_{nj}(x)}\, M_{nm}(x) g_{mk}(x)\, d\mu_\tr(x).
465: \end{equation}
466: Since $f^\dagger Mf\geq 0$, we see that
467: \begin{equation} \lb{1.25a}
468: \ang{f,f}_R \geq 0.
469: \end{equation}
470: One might be tempted to think of $\ang{f,f}_R^{1/2}$ as some kind of
471: norm, but that is
472: doubtful. Even if $\mu$ is supported at a single point, $x_0$, with
473: $M=l^{-1}\bdone$, this
474: ``norm'' is essentially the absolute value of $A=f(x_0)$, which is
475: known not to obey the
476: triangle inequality! (See \cite[Sect.~I.1]{S-TI} for an example.)
477: 
478: However, if one looks at
479: \begin{equation} \lb{1.26}
480: \norm{f}_R=(\Tr\ang{f,f}_R)^{1/2},
481: \end{equation}
482: one does have a norm (or, at least, a semi-norm). Indeed,
483: \begin{equation} \lb{1.27}
484: \langle f,g\rangle_R=\Tr\ang{f,g}_R
485: \end{equation}
486: is a sesquilinear form which is positive semi-definite, so
487: \eqref{1.26} is the semi-norm corresponding to an inner product
488: and, of course, one has a Cauchy--Schwarz inequality
489: \begin{equation} \lb{1.28}
490: \abs{\Tr\ang{f,g}_R}\leq \norm{f}_R \norm{g}_R.
491: \end{equation}
492: 
493: We have not specified which $f$'s and $g$'s can be used in \eqref
494: {1.24}. We have in mind
495: mainly polynomials in $x$ in the real case and Laurent polynomials in
496: $z$ in the $\partial\bbD$
497: case although, obviously, continuous functions are okay. Indeed, it
498: suffices that $f$ (and $g$) be measurable and
499: obey
500: \begin{equation} \lb{1.28a}
501: \int \Tr(f^\dagger (x) f(x))\, d\mu_\tr (x) <\infty
502: \end{equation}
503: for the integrals in \eqref{1.25} to converge. The set of
504: equivalence classes under $f\sim g$ if $\norm{f-g}_R = 0$ defines
505: a Hilbert space, $\calH$, and $\langle f,g\rangle_R$ is
506: the inner product on this space.
507: 
508: Instead of \eqref{1.24}, we use the suggestive shorthand
509: \begin{equation} \lb{1.29}
510: \ang{f,g}_R = \int f(x)^\dagger \, d\mu(x) g(x).
511: \end{equation}
512: The use of $R$ here comes from ``right'' for if $\alpha\in\calM_l$,
513: \begin{align}
514: \ang{f,g\alpha}_R &= \ang{f,g}_R \alpha, \lb{1.30} \\
515: \ang{f\alpha,g}_R &= \alpha^\dagger\ang{f,g}_R, \lb{1.31}
516: \end{align}
517: but, in general, $\ang{f,\alpha g}_R$ is not related to $\ang{f,g}_R$.
518: 
519: While $(\Tr\ang{f,f}_R)^{1/2}$ is a natural analogue of the norm
520: in the scalar case, it will sometimes be useful to instead
521: consider
522: \begin{equation} \lb{1.31a}
523: [\det \ang{f,f}_R]^{1/2}.
524: \end{equation}
525: Indeed, this is a stronger ``norm'' in that $\det >0\Rightarrow
526: \Tr>0$ but not vice-versa.
527: 
528: When $d\mu$ is a ``direct sum,'' that is, each $M(x)$ is diagonal,
529: one can appreciate the difference.
530: In that case, $d\mu=d\mu_1\oplus\cdots\oplus d\mu_l$ and the MOPRL
531: are direct sums (i.e., diagonal
532: matrices) of scalar OPRL
533: \begin{equation} \lb{1.31b}
534: P_n^R (x,d\mu) =P_n(x,d\mu_1) \oplus\cdots\oplus P_n(x,d\mu_l).
535: \end{equation}
536: Then
537: \begin{equation} \lb{1.31c}
538: \norm{P_n^R}_R =\biggl(\, \sum_{j=1}^l \norm{P_n(\cdot, d\mu_j)}_{L^2
539: (d\mu_j)}^2\biggr)^{1/2},
540: \end{equation}
541: while
542: \begin{equation} \lb{1.31d}
543: (\det \ang{P_n^R,P_n^R}_R)^{1/2} = \prod_{j=1}^l \norm{P_n (\cdot, d
544: \mu_j)}_{L^2(d\mu_j)}.
545: \end{equation}
546: In particular, in terms of extending the theory of regular
547: measures \cite{StT}, $\norm{P_n^R}^{1/n}_R$ is only sensitive to
548: $\max\norm{P_n (\cdot, d\mu_j)}_{L^2(d\mu_j)}^{1/2} $ while
549: $(\det\ang{P_n^R,P_n^R}_R)^{1/2}$ is sensitive to them all. Thus,
550: $\det$ will be needed for that theory (see Chapter~\ref{s5}).
551: 
552: There will also be a left inner product and, correspondingly, two
553: sets of MOPRL and MOPUC. We discuss
554: this further in Sections~\ref{s2.1} and \ref{s3.1}.
555: 
556: Occasionally, for $\bbC^l$ vector-valued functions $f$ and $g$, we
557: will want to consider the scalar
558: \begin{equation} \lb{1.32}
559: \sum_{k,j}\int \ol{f_k(x)}\, M_{kj}(x) g_j (x)\, d\mu_\tr(x),
560: \end{equation}
561: which we will denote
562: \begin{equation} \lb{1.33}
563: \int d \langle f(x),\mu(x) g(x)\rangle_{\C^l}.
564: \end{equation}
565: 
566: 
567: We next turn to module Fourier expansions. A set $\{\varphi_j\}_{j=1}
568: ^N$ in $\calH$ ($N$ may be infinite)
569: is called orthonormal if and only if
570: \begin{equation} \lb{1.33a}
571: \ang{\varphi_j,\varphi_k}_R =\delta_{jk}\bdone.
572: \end{equation}
573: This natural terminology is an abuse of notation since
574: \eqref{1.33a} implies orthogonality in $\langle\cdot,\cdot
575: \rangle_R$ but not normalization, and is much stronger than orthogonality in
576: $\langle \cdot,\cdot\rangle_R$.
577: 
578: Suppose for a moment that $N<\infty$. For any $a_1, \dots, a_N\in
579: \calM_l$, we can form $\sum_{j=1}^N
580: \varphi_j a_j$ and, by the right multiplication relations \eqref
581: {1.30}, \eqref{1.31}, and \eqref{1.33a},
582: we have
583: \begin{equation} \lb{1.33b}
584: \biggl< \!\!\!\biggl< \sum_{j=1}^N \varphi_j a_j, \sum_{j=1}^N
585: \varphi_j b_j \biggr>\!\! \!\biggr>_R =
586: \sum_{j=1}^N a_j^\dagger b_j.
587: \end{equation}
588: We will denote the set of all such $\sum_{j=1}^N \varphi_j a_j$ by $
589: \calH_{(\varphi_j)}$---it is a
590: vector subspace of $\calH$ of dimension (over $\bbC$) $Nl^2$.
591: 
592: Define for $f\in\calH$,
593: \begin{equation} \lb{1.33c}
594: \pi_{(\varphi_j)} (f)=\sum_{j=1}^N \varphi_j \ang{\varphi_j,f}_R.
595: \end{equation}
596: It is easy to see it is the orthogonal projection in the scalar
597: inner product $\langle \cdot,\cdot\rangle_R$ from $\calH$ to $\calH_{(\varphi_j)}$.
598: 
599: By the standard Hilbert space calculation, taking care to only
600: multiply on the right, one finds the
601: Pythagorean theorem,
602: \begin{equation} \lb{1.33d}
603: \ang{f,f}_R = \ang{f-\pi_{(\varphi_j)} f, f-\pi_{(\varphi_j)}f}_R +
604: \sum_{j=1}^N
605: \ang{\varphi_j,f}_R^\dagger \ang{\varphi_j,f}_R.
606: \end{equation}
607: 
608: As usual, this proves for infinite $N$ that
609: \begin{equation} \lb{1.33e}
610: \sum_{j=1}^N \ang{\varphi_j,f}_R^\dagger \ang{\varphi_j,f}_R \leq \ang
611: {f,f}_R
612: \end{equation}
613: and the convergence of
614: \begin{equation} \lb{1.33f}
615: \sum_{j=1}^N \varphi_j \ang{\varphi_j,f}_R \equiv \pi_{(\varphi_j)}(f)
616: \end{equation}
617: allowing the definition of $\pi_{(\varphi_j)}$ and of $\calH_
618: {(\varphi_j)}\equiv\Ran \pi_{(\varphi_j)}$
619: for $N=\infty$.
620: 
621: An orthonormal set is called complete if
622: $\calH_{(\varphi_j)}=\calH$. In that case, equality holds in
623: \eqref{1.33e} and $\pi_{(\varphi_j)}(f)=f$.
624: 
625: For orthonormal bases, we have the Parseval relation from
626: \eqref{1.33d}
627: \begin{equation} \lb{1.33g}
628: \ang{f,f}_R =\sum_{j=1}^\infty \ang{\varphi_j,f}_R^\dagger \ang
629: {\varphi_j,f}_R
630: \end{equation}
631: and
632: \begin{equation}\lb{1.33h}
633: \norm{f}_R^2 =\sum_{j=1}^\infty \Tr(\ang{\varphi_j,f}_R^\dagger
634: \ang {\varphi_j,f}_R).
635: \end{equation}
636: 
637: \subsection{Matrix M\"obius Transformations} \lb{s1.3}
638: 
639: Without an understanding of matrix M\"obius transformations, the
640: form of the MOPUC Geronimus theorem we will prove in
641: Section~\ref{s3.5} will seem strange-looking. To set the stage,
642: recall that scalar fractional linear transformations (FLT) are
643: associated to matrices $T=\left( \begin{smallmatrix} a&b \\ c&d
644: \end{smallmatrix}\right)$ with $\det T \not= 0$ via
645: \begin{equation} \lb{1.20x}
646: f_T(z) = \f{az+b}{cz+d}\, .
647: \end{equation}
648: Without loss, one can restrict to
649: \begin{equation} \lb{1.21x}
650: \det(T)=1.
651: \end{equation}
652: Indeed, $T\mapsto f_T$ is a $2$ to $1$ map of $\bbS\bbL(2,\bbC)$ to
653: maps of $\bbC\cup\{\infty\}$ to itself.
654: One advantage of the matrix formalism is that the map is a matrix
655: homomorphism, that is,
656: \begin{equation} \lb{1.22x}
657: f_{T\circ S} = f_T \circ f_S,
658: \end{equation}
659: which shows that the group of FLTs is $\bbS\bbL(2,\bbC)/\{\bdone, -
660: \bdone\}$.
661: 
662: While \eqref{1.22x} can be checked by direct calculation, a more
663: instructive way is to look at the
664: complex projective line. $u,v\in\bbC^2\setminus\{0\}$ are called
665: equivalent if there is $\lambda\in
666: \bbC\setminus\{0\}$ so that $u=\lambda v$. Let $[\cdot]$ denote
667: equivalence classes. Except for
668: $[\binom{1}{0}]$, every equivalence class contains exactly one point
669: of the form $\binom{z}{1}$ with
670: $z\in\bbC$. If $[\binom{1}{0}]$ is associated with $\infty$, the set
671: of equivalence classes is naturally
672: associated with $\bbC\cup\{\infty\}$. $f_T$ then obeys
673: \begin{equation} \lb{1.23x}
674: \biggl[ T\binom{z}{1}\biggr] = \biggl[ \binom{f_T(z)}{1}\biggr]
675: \end{equation}
676: from which \eqref{1.22x} is immediate.
677: 
678: By M\"obius transformations we will mean those FLTs that map
679: $\bbD$ onto itself. Let
680: \begin{equation} \lb{1.24x}
681: J=\left( \begin{array}{rr}
682: 1 & 0 \\
683: 0 & -1
684: \end{array} \right).
685: \end{equation}
686: Then $[u]=[\binom{z}{1}]$ with $\abs{z}=1$ (resp.\ $\abs{z}<1$) if
687: and only if $\jap{u,Ju}=0$
688: (resp.\ $\jap{u,Ju}<0$). From this, it is not hard to show that if $
689: \det(T)=1$, then $f_T$ maps
690: $\bbD$ invertibly onto $\bbD$ if and only if
691: \begin{equation} \lb{1.25x}
692: T^\dagger \! JT=J.
693: \end{equation}
694: If $T$ has the form $\left(\begin{smallmatrix} a&b \\ c&d \end
695: {smallmatrix}\right)$, this is equivalent to
696: \begin{equation} \lb{1.26x}
697: \abs{a}^2 - \abs{c}^2 =1, \qquad \abs{b}^2-\abs{d}^2 =-1, \qquad \bar
698: ab-\bar cd =0.
699: \end{equation}
700: The set of $T$'s obeying $\det(T)=1$ and \eqref{1.25x} is called $\bbS
701: \bbU(1,1)$. It is
702: studied extensively in \cite[Sect.~10.4]{S2}.
703: 
704: The self-adjoint elements of $\bbS\bbU(1,1)$ are parametrized by $
705: \alpha\in\bbD$ via $\rho =
706: (1-\abs{\alpha}^2)^{1/2}$,
707: \begin{equation} \lb{1.27x}
708: T_\alpha = \f{1}{\rho}\begin{pmatrix}
709: 1 & \alpha \\
710: \bar\alpha & 1 \end{pmatrix}
711: \end{equation}
712: associated to
713: \begin{equation} \lb{1.28x}
714: f_{T_\alpha}(z) = \f{z+\alpha}{1+\bar\alpha z}\, .
715: \end{equation}
716: Notice that
717: \begin{equation} \lb{1.29x}
718: T_\alpha^{-1} = T_{-\alpha}
719: \end{equation}
720: and that
721: \[
722: \forall z\in\bbD,\, \exists \, ! \, \alpha \text{ such that } T_
723: \alpha (0)=z,
724: \]
725: namely, $\alpha =z$.
726: 
727: It is a basic theorem that every holomorphic bijection of $\bbD$ to $
728: \bbD$ is an $f_T$ for some
729: $T$ in $\bbS\bbU(1,1)$ (unique up to $\pm\bdone$).
730: 
731: With this in place, we can turn to the matrix case. Let $\calM_l$ be
732: the space of $l\times l$
733: complex matrices with the Euclidean norm induced by the vector norm
734: $\jap{\cdot, \cdot}_{\bbC^l}^{1/2}$. Let
735: \begin{equation} \lb{1.30x}
736: \bbD_l =\{A\in\calM_l\, \colon \norm{A} <1\}.
737: \end{equation}
738: We are interested in holomorphic bijections of $\bbD_l$ to itself,
739: especially via a suitable notion of
740: FLT. There is a huge (and diffuse) literature on the subject,
741: starting with its use in analytic number
742: theory. It has also been studied in connection with electrical
743: engineering filters and indefinite
744: matrix Hilbert spaces. Among the huge literature, we mention \cite
745: {ALK,Alp,Dym,GK,Helg,ScZ}. Especially
746: relevant to MOPUC is the book of Bakonyi--Constantinescu \cite{BakCon}.
747: 
748: Consider $\calM_l \oplus \calM_l = \calM_l [2]$ as a right module
749: over $\calM_l$. The $\calM_l$-projective line is defined by saying
750: $\left[\begin{smallmatrix} X \\ Y
751: \end{smallmatrix}\right]\sim \left[\begin{smallmatrix} X' \\
752: Y'\end{smallmatrix}\right]$, both in $\calM_l[2]
753: \setminus\{\bdzero\}$, if and only if there exists
754: $\Lambda\in\calM_l$, $\Lambda$ invertible so that
755: \begin{equation} \lb{1.31x}
756: X= X'\Lambda, \qquad Y=Y'\Lambda.
757: \end{equation}
758: Let $T$ be a map of $\calM_l[2]$ of the form
759: \begin{equation} \lb{1.32x}
760: T=\begin{pmatrix} A&B \\ C&D
761: \end{pmatrix}
762: \end{equation}
763: acting on $\calM_l[2]$ by
764: \begin{equation} \lb{1.33x}
765: T\begin{bmatrix} X \\ Y \end{bmatrix} =
766: \begin{bmatrix} AX + BY \\ CX + DY \end{bmatrix}.
767: \end{equation}
768: Because this acts on the left and $\Lambda$ equivalence on the right,
769: $T$ maps equivalence classes
770: to themselves. In particular, if $CX+D$ is invertible, $T$ maps the
771: equivalence class of
772: $\left[\begin{smallmatrix}X \\ \bdone
773: \end{smallmatrix}\right]$ to the equivalence class of
774: $\left[\begin{smallmatrix}f_T[X] \\ \bdone \end{smallmatrix}\right]$,
775: where
776: \begin{equation} \lb{1.34}
777: f_T[X]=(AX+B)(CX+D)^{-1}.
778: \end{equation}
779: 
780: So long as $CX+D$ remains invertible, \eqref{1.22x} remains true. Let
781: $J$ be the $2l\times 2l$ matrix
782: in $l\times l$ block form
783: \begin{equation} \lb{1.35}
784: J = \left( \begin{array}{rr}
785: \bdone  & \bdzero \\ \bdzero & -\bdone
786: \end{array} \right).
787: \end{equation}
788: Note that (with $\left[\begin{smallmatrix} X \\ \bdone \end
789: {smallmatrix}\right]^\dagger = [X^\dagger \bdone]$)
790: \begin{equation} \lb{1.36}
791: \begin{bmatrix} X \\ \bdone \end{bmatrix}^\dagger J \begin{bmatrix} X
792: \\ \bdone \end{bmatrix} \leq
793: \bdzero \Leftrightarrow X^\dagger X\leq \bdone \Leftrightarrow \norm{X}
794: \leq \bdone.
795: \end{equation}
796: Therefore, if we define $\bbS\bbU(l,l)$ to be those $T$'s with
797: $\det T=1$ and
798: \begin{equation} \lb{1.37}
799: T^\dagger \! JT =J,
800: \end{equation}
801: then
802: \begin{equation} \lb{1.38}
803: T\in\bbS\bbU(l,l) \Rightarrow f_T [\bbD_l]=\bbD_l \text{ as a
804: bijection}.
805: \end{equation}
806: If $T$ has the form \eqref{1.32x}, then \eqref{1.37} is equivalent to
807: \begin{align}
808: & A^\dagger A - C^\dagger C = D^\dagger D - B^\dagger B =\bdone, \lb
809: {1.61a} \\
810: & A^\dagger B = C^\dagger D \lb{1.61b}
811: \end{align}
812: (the fourth relation $B^\dagger A = D^\dagger C$ is equivalent to
813: \eqref{1.61b}).
814: 
815: This depends on
816: 
817: \begin{proposition}\lb{P1.3.1} If $T=\left(\begin{smallmatrix} A&B \\
818: C&D\end{smallmatrix}\right)$ obeys \eqref{1.37} and $\norm{X} <1$,
819: then $CX+D$ is invertible.
820: \end{proposition}
821: 
822: \begin{proof} \eqref{1.37} implies that
823: \begin{align}
824: T^{-1} &= JT^\dagger \! J \lb{1.37a} \\
825: &= \left( \begin{array}{rr}
826: A^\dagger & -C^\dagger \\
827: -B^\dagger & D^\dagger
828: \end{array} \right). \lb{1.37b}
829: \end{align}
830: Clearly, \eqref{1.37} also implies $T^{-1}\in\bbS\bbU(l,l)$. Thus,
831: by \eqref{1.61a} for $T^{-1}$,
832: \begin{equation} \lb{1.37c}
833: DD^\dagger -CC^\dagger =\bdone.
834: \end{equation}
835: This implies first that $DD^\dagger\geq \bdone$, so $D$ is
836: invertible, and second that
837: \begin{equation} \lb{1.37d}
838: \norm{D^{-1} C}\leq 1.
839: \end{equation}
840: Thus, $\norm{X}<1$ implies $\norm{D^{-1}CX}<1$ so $\bdone+D^{-1}CX$
841: is invertible, and thus
842: so is $D(\bdone+D^{-1}CX)$.
843: \end{proof}
844: 
845: It is a basic result of Cartan \cite{Cartan} (see Helgason
846: \cite{Helg} and the discussion therein) that
847: 
848: \begin{theorem}\lb{T1.3.2} A holomorphic bijection, $g$, of $\bbD_l$
849: to itself is either of
850: the form
851: \begin{equation} \lb{1.38x}
852: g(X)=f_T(X)
853: \end{equation}
854: for some $T\in\bbS\bbU(l,l)$ or
855: \begin{equation} \lb{1.39}
856: g(X)=f_T(X^t).
857: \end{equation}
858: \end{theorem}
859: 
860: Given $\alpha\in\calM_l$ with $\norm{\alpha}<1$, define
861: \begin{equation} \lb{1.40}
862: \rho^L = (\bdone-\alpha^\dagger \alpha)^{1/2}, \qquad
863: \rho^R =(\bdone-\alpha\alpha^\dagger)^{1/2}.
864: \end{equation}
865: 
866: \begin{lemma} \lb{L1.3.3} We have
867: \begin{alignat}{2}
868: \alpha\rho^L &= \rho^R\alpha, \qquad & \alpha^\dagger \rho^R &=
869: \rho^L \alpha^\dagger, \lb{1.41} \\
870: \alpha(\rho^L)^{-1} &= (\rho^R)^{-1}\alpha, \qquad & \alpha^\dagger
871: (\rho^R)^{-1} &= (\rho^L)^{-1}
872: \alpha^\dagger. \lb{1.42}
873: \end{alignat}
874: \end{lemma}
875: 
876: \begin{proof} Let $f$ be analytic in $\bbD$ with $f(z)=\sum_{n=0}^
877: \infty c_n z^n$ its Taylor series at $z=0$.
878: Since $\norm{\alpha^\dagger\alpha}<1$, we have
879: \begin{equation} \lb{1.43}
880: f(\alpha^\dagger \alpha) = \sum_{n=0}^\infty c_n (\alpha^\dagger
881: \alpha)^n
882: \end{equation}
883: norm convergent, so $\alpha (\alpha^\dagger\alpha)^n = (\alpha\alpha^
884: \dagger)^n\alpha$ implies
885: \begin{equation} \lb{1.44}
886: \alpha f(\alpha^\dagger\alpha) = f(\alpha\alpha^\dagger)\alpha,
887: \end{equation}
888: which implies the first halves of \eqref{1.41} and \eqref{1.42}. The
889: other halves follow by
890: taking adjoints.
891: \end{proof}
892: 
893: \begin{theorem}\lb{T1.4} There is a one-one correspondence between $
894: \alpha$'s in $\calM_l$ obeying
895: $\norm{\alpha}<1$ and positive self-adjoint elements of $\bbS\bbU(l,l)
896: $ via
897: \begin{equation} \lb{1.45}
898: T_\alpha = \begin{pmatrix}
899: (\rho^R)^{-1} & (\rho^R)^{-1}\alpha \\
900: (\rho^L)^{-1} \alpha^\dagger & (\rho^L)^{-1}
901: \end{pmatrix}.
902: \end{equation}
903: \end{theorem}
904: 
905: \begin{proof} A straightforward calculation using Lemma~\ref{L1.3.3}
906: proves that $T_\alpha$ is self-adjoint and $T_\alpha^\dagger
907: JT_\alpha =J$. Conversely, if $T$ is self-adjoint, $T=\left(
908: \begin{smallmatrix} A&B\\C&D\end{smallmatrix} \right)$ and in
909: $\bbS\bbU(l,l)$, then $T^\dagger = T\Rightarrow A^\dagger = A$,
910: $B^\dagger =C$, so \eqref{1.61a} becomes
911: \begin{equation} \lb{1.46}
912: AA^\dagger - BB^\dagger =\bdone,
913: \end{equation}
914: so if
915: \begin{equation} \lb{1.47}
916: \alpha = A^{-1} B,
917: \end{equation}
918: then \eqref{1.46} becomes
919: \begin{equation} \lb{1.48}
920: A^{-1} (A^{-1})^\dagger + \alpha\alpha^\dagger =\bdone.
921: \end{equation}
922: Since $T\geq 0$, $A\geq 0$ so \eqref{1.48} implies $A=(\rho^R)^{-1}$,
923: and then \eqref{1.47} implies
924: $B=(\rho^R)^{-1} \alpha$.
925: 
926: By Lemma~\ref{L1.3.3},
927: \begin{equation} \lb{1.49}
928: C=B^\dagger = \alpha^\dagger (\rho^R)^{-1} = (\rho^L)^{-1} \alpha^
929: \dagger
930: \end{equation}
931: and then (by $D=D^\dagger$, $C^\dagger =B$, and \eqref{1.61a})
932: $D D^\dagger - C C^\dagger =\bdone$ plus $D>0$ implies
933: $D= (\rho^L)^{-1}$.
934: \end{proof}
935: 
936: \begin{corollary}\lb{C1.3.5} For each $\alpha\in\bbD_l$, the map
937: \begin{equation} \lb{1.50}
938: f_{T_\alpha}(X) = (\rho^R)^{-1} (X+\alpha) (\bdone+\alpha^\dagger X)^
939: {-1} (\rho^L)
940: \end{equation}
941: takes $\bbD_l$ to $\bbD_l$. Its inverse is given by
942: \begin{equation} \lb{1.51}
943: f_{T_\alpha}^{-1}(X) = f_{T_{-\alpha}}(X) = (\rho^R)^{-1} (X-\alpha)
944: (\bdone-\alpha^\dagger X)^{-1} (\rho^L).
945: \end{equation}
946: \end{corollary}
947: 
948: There is an alternate form for the right side of \eqref{1.50}.
949: 
950: \begin{proposition} \lb{P1.6} The following identity holds true for any $X$,
951: $\norm{X}\leq 1$:
952: \begin{equation} \lb{1.82a}
953: \rho^R (1+X\alpha^\dagger)^{-1} (X+\alpha)(\rho^L)^{-1} =
954: (\rho^R)^{-1} (X+\alpha)(1+\alpha^\dagger X)^{-1}\rho^L.
955: \end{equation}
956: \end{proposition}
957: 
958: \begin{proof}
959: By the definition of $\rho^L$ and $\rho^R$, we have
960: $$
961: X (\rho^L)^{-2}(1-\alpha^\dagger\alpha)=(\rho^R)^{-2}(1-\alpha\alpha^\dagger)X.
962: $$
963: Expanding, using \eqref{1.42} and rearranging, we get
964: $$
965: X(\rho^L)^{-2}+\alpha(\rho^L)^{-2}\alpha^\dagger X
966: =
967: (\rho^R)^{-2}X+X\alpha^\dagger(\rho^R)^{-2}\alpha.
968: $$
969: Adding $\alpha(\rho^L)^{-2}+X(\rho^L)^{-2}\alpha^\dagger X$ to both sides and
970: using \eqref{1.42} again, we obtain
971: \begin{multline*}
972: X(\rho^L)^{-2}+\alpha(\rho^L)^{-2}+X(\rho^L)^{-2}\alpha^\dagger X
973: +\alpha(\rho^L)^{-2}\alpha^\dagger X
974: \\
975: =
976: (\rho^R)^{-2}X+(\rho^R)^{-2}\alpha+X\alpha^\dagger(\rho^R)^{-2}X
977: +X\alpha^\dagger(\rho^R)^{-2}\alpha,
978: \end{multline*}
979: which is the same as
980: $$
981: (X+\alpha)(\rho^L)^{-2}(1+\alpha^\dagger X)
982: =
983: (1+X\alpha^\dagger)(\rho^R)^{-2}(X+\alpha).
984: $$
985: Multiplying by $(1+X\alpha^\dagger)^{-1}$ and $(1+\alpha^\dagger X)^{-1}$, we get
986: $$
987: (1+X\alpha^\dagger)^{-1}(X+\alpha)(\rho^L)^{-2}
988: =
989: (\rho^R)^{-2}(X+\alpha)(1+\alpha^\dagger X)^{-1}
990: $$
991: and the statement follows.
992: \end{proof}
993: 
994: 
995: \subsection{Applications and Examples} \lb{s1.4}
996: 
997: There are a number of simple examples which show that beyond their
998: intrinsic mathematical interest, MOPRL
999: and MOPUC have wide application.
1000: 
1001: \subsubsection*{{\rm (a)} Jacobi matrices on a strip} Let $\Lambda
1002: \subset\bbZ^\nu$ be a subset (perhaps infinite) of the
1003: $\nu$-dimensional lattice $\bbZ^\nu$ and let $\ell^2 (\Lambda)$ be
1004: square summable sequences indexed by $\Lambda$. Suppose a real
1005: symmetric matrix $\alpha_{ij}$ is given for all $i,j\in\Lambda$
1006: with $\alpha_{ij}=0$ unless $\abs{i-j}=1$ (nearest neighbors). Let
1007: $\beta_i$ be a real sequence indexed by $i\in\Lambda$. Suppose
1008: \begin{equation} \lb{1.40x}
1009: \sup_{i,j}\, \abs{\alpha_{ij}} + \sup_i\, \abs{\beta_i} <\infty.
1010: \end{equation}
1011: Define a bounded operator, $J$, on $\ell^2(\Lambda)$ by
1012: \begin{equation} \lb{1.41x}
1013: (Ju)_i = \sum_j \alpha_{ij} u_j + \beta_i u_i.
1014: \end{equation}
1015: The sum is finite with at most $2\nu$ elements.
1016: 
1017: The special case $\Lambda =\{1,2,\dots\}$ with $b_i =\beta_i$, $a_i =
1018: \alpha_{i,i+1}>0$
1019: corresponds precisely to classical semi-infinite tridiagonal Jacobi
1020: matrices.
1021: 
1022: Now consider the situation where $\Lambda' \subset\bbZ^{\nu-1}$ is a
1023: finite set with $l$ elements
1024: and
1025: \begin{equation} \lb{1.42x}
1026: \Lambda = \{j\in\bbZ^\nu \, \colon j_1\in \{1,2,\dots\};\, (j_2,
1027: \dots j_\nu)\in\Lambda'\},
1028: \end{equation}
1029: a ``strip'' with cross-section $\Lambda'$. $J$ then has a block $l
1030: \times l$ matrix Jacobi form
1031: where $(\gamma,\delta\in\Lambda'$)
1032: \begin{alignat}{2}
1033: (B_i)_{\gamma\delta} &= b_{(i,\gamma)}, && \qquad (\gamma=\delta), \lb
1034: {1.42y} \\
1035: &= a_{(i,\gamma)(i,\delta)}, && \qquad (\gamma\neq\delta), \lb{1.42z} \\
1036: (A_i)_{\gamma\delta} &= a_{(i,\gamma)(i+1,\delta)}. \lb{1.43x}
1037: \end{alignat}
1038: The nearest neighbor condition says $(A_i)_{\gamma\delta}=0$ if $
1039: \gamma\neq\delta$. If
1040: \begin{equation} \lb{1.44x}
1041: a_{(i,\gamma)(i+1,\gamma)} >0
1042: \end{equation}
1043: for all $i,\gamma$, then $A_i$ is invertible and we have a block
1044: Jacobi matrix of the
1045: kind described in Section~\ref{s2.2} below.
1046: 
1047: By allowing general $A_i,B_i$, we obtain an obvious generalization of
1048: this model---an interpretation of
1049: general MOPRL.
1050: 
1051: Schr\"odinger operators on strips have been studied in part as
1052: approximations to $\bbZ^\nu$; see \cite{CrS,GKM,KS88,Lac,MV,SB04}.
1053: {}From this point of view, it is also natural to allow periodic
1054: boundary conditions in the vertical directions. Furthermore, there
1055: is closely related work on Schr\"odinger (and other) operators
1056: with matrix-valued potentials; see, for example, \cite{BGMS03,
1057: CG01,CG02,CG03,CG04, CGHL00, CGZ07,GS03, GS00, SB07}.
1058: 
1059: \subsubsection*{{\rm (b)} Two-sided Jacobi matrices} This example goes
1060: back at least to Nikishin \cite{Nik}. Consider the
1061: case $\nu=2$, $\Lambda'=\{0,1\}
1062: \subset\bbZ$, and $\Lambda$ as above. Suppose \eqref{1.44x} holds,
1063: and in addition,
1064: \begin{align}
1065: a_{(1,0)(1,1)} &>0, \lb{1.45x} \\
1066: a_{(i,0)(i,1)} &=0,  \quad i=2,3,\dots . \lb{1.46x}
1067: \end{align}
1068: Then there are no links between the rungs of the ``ladder,'' $\{1,2,
1069: \dots\}\times \{0,1\}$ except at the
1070: end and the ladder can be unfolded to $\bbZ$! Thus, a two-sided
1071: Jacobi matrix can be viewed as a special
1072: kind of one-sided $2\times 2$ matrix Jacobi operator.
1073: 
1074: It is known that for two-sided Jacobi matrices, the spectral theory
1075: is determined by the $2\times 2$
1076: matrix
1077: \begin{equation} \lb{1.47x}
1078: d\mu = \begin{pmatrix}
1079: d\mu_{00} & d\mu_{01} \\
1080: d\mu_{10} & d\mu_{11}
1081: \end{pmatrix},
1082: \end{equation}
1083: where $d\mu_{kl}$ is the measure with
1084: \begin{equation} \lb{1.48x}
1085: \jap{\delta_k, (J-\lambda)^{-1}\delta_l} = \int \f{d\mu_{kl}(x)}{x-
1086: \lambda},
1087: \end{equation}
1088: but also that it is very difficult to specify exactly which $d\mu$
1089: correspond to two-sided Jacobi
1090: matrices.
1091: 
1092: This difficulty is illuminated by the theory of MOPRL. By Favard's
1093: theorem (see Theorem~\ref{favard}),
1094: every such $d\mu$ (given by \eqref{1.47x} and positive definite and
1095: non-trivial in a sense we will
1096: describe in Lemma~\ref{nondeg}) yields a unique block Jacobi matrix
1097: with $A_j>0$ (positive definite).
1098: This $d\mu$ comes from a two-sided Jacobi matrix if and only if
1099: \begin{SL}
1100: \item[(a)] $B_j$ is diagonal for $j=2,3,\dots$.
1101: \item[(b)] $A_j$ is diagonal for $j=1,2,\dots$.
1102: \item[(c)] $B_j$ has strictly positive off-diagonal elements.
1103: \end{SL}
1104: These are very complicated indirect conditions on $d\mu$!
1105: 
1106: \subsubsection*{{\rm (c)} Banded matrices} Classical Jacobi matrices
1107: are semi-infinite symmetric
1108: tridiagonal matrices, that is,
1109: \begin{equation} \lb{1.49x}
1110: J_{km}=0 \quad\text{if } \abs{k-m}>1
1111: \end{equation}
1112: with
1113: \begin{equation} \lb{1.50x}
1114: J_{km}>0 \quad\text{if } \abs{k-m}=1.
1115: \end{equation}
1116: 
1117: A natural generalization are $(2l+1)$-diagonal symmetric matrices,
1118: that is,
1119: \begin{alignat}{2}
1120: J_{km} &= 0 &  \quad \text{if } \abs{k-m} &>l \lb{1.51x},  \\
1121: J_{km} &> 0 & \quad \text{if } \abs{k-m} &=l. \lb{1.52x}
1122: \end{alignat}
1123: 
1124: Such a matrix can be partitioned into $l\times l$ blocks, which is
1125: tridiagonal in block. The
1126: conditions \eqref{1.51x} and \eqref{1.52x} are equivalent to $A_k\in
1127: \calL$, the set of lower
1128: triangular matrices; and conversely, $A_k\in\calL$, with $A_k,B_k$
1129: real (and $B_k$ symmetric)
1130: correspond precisely to such banded matrices. This is why we
1131: introduce type~3 MOPRL.
1132: 
1133: Banded matrices correspond to certain higher-order difference
1134: equations. Unlike the second-order
1135: equation (which leads to tridiagonal matrices) where every equation
1136: with positive coefficients is
1137: equivalent via a variable rescaling to a symmetric matrix, only
1138: certain higher-order difference
1139: equations correspond to symmetric block Jacobi matrices.
1140: 
1141: \subsubsection*{{\rm (d)} Magic formula} In \cite{DKSppt}, Damanik,
1142: Killip, and Simon studied
1143: perturbations of Jacobi and CMV matrices with periodic Jacobi
1144: parameters (or Verblunsky coefficients).
1145: They proved that if $\Delta$ is the discriminant of a two-sided
1146: periodic $J_0$, then a bounded
1147: two-sided $J$ has $\Delta(J)=S^p + S^{-p}$ ($(Su)_n\equiv u_{n+1}$)
1148: if and only if $J$ lies
1149: in the isospectral torus of $J_0$. They call this the magic formula.
1150: 
1151: This allows the study of perturbations of the isospectral torus by
1152: studying $\Delta(J)$ which is
1153: a polynomial in $J$ of degree $p$, and so a $2p+1$ banded matrix.
1154: Thus, the study of perturbations
1155: of periodic problems is connected to perturbations of $S^p + S^{-p}$
1156: as block Jacobi matrices.
1157: Indeed, it was this connection that stimulated our interest in MOPRL,
1158: and \cite{DKSppt} uses some
1159: of our results here.
1160: 
1161: \subsubsection*{{\rm (e)} Vector-valued prediction theory} As noted
1162: in Section~\ref{s1.1},
1163: both prediction theory and filtering theory use OPUC and have natural
1164: MOPUC settings that
1165: motivated much of the MOPUC literature.
1166: 
1167: 
1168: 
1169: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
1170: \section{Matrix Orthogonal Polynomials on the Real Line} \lb{s2}
1171: 
1172: \subsection{Preliminaries} \lb{s2.1}
1173: 
1174: OPRL are the most basic and developed of orthogonal polynomials,
1175: and so this chapter on the matrix analogue is the most important
1176: of this survey. We present the basic formulas, assuming enough
1177: familiarity with the scalar case (see
1178: \cite{Chi,FrB,S,Rice,Szb,Teschl}) that we do not need to explain
1179: why the objects we define are important.
1180: 
1181: \subsubsection{Polynomials, Inner Products, Norms} \lb{s2.1.1}
1182: 
1183: Let $d\mu$ be an $l\times l$ matrix-valued Hermitian positive
1184: semi-definite finite measure on $\R$
1185: normalized by $\mu(\R)={\boldsymbol 1}\in\calM_l $.
1186: We assume for simplicity that $\mu$ has a compact support.
1187: However, many of the results below do not need the latter restriction and in fact can be
1188: found in the literature for matrix-valued measures with unbounded support.
1189: 
1190: 
1191: Define (as in
1192: \eqref{1.24})
1193: \begin{alignat*}{3}
1194: \ang{f , g}_R &= \int f(x)^\dagger \, d\mu(x) \, g(x), \quad & \norm
1195: {f}_R & =(\Tr \ang{f,f}_R)^{1/2},
1196: \quad & f,g & \in\pol, \\
1197: \ang{f , g}_L &= \int g(x)\, d\mu(x) \, f(x)^\dagger,
1198: \quad & \norm{f}_L &=(\Tr \ang{f,f}_L)^{1/2}, \quad & f,g &\in \pol.
1199: \end{alignat*}
1200: Clearly, we have
1201: \begin{alignat}{2}
1202: \ang{f,g}_R^\dagger &=\ang{g,f}_R, \qquad
1203: & \ang{f,g}_L^\dagger & =\ang{g,f}_L, \label{b1} \\
1204: \ang{f,g}_L &=\ang{g^\dagger,f^\dagger}_R, \qquad
1205: & \norm{f}_L &=\norm{f^\dagger}_R. \label{b2}
1206: \end{alignat}
1207: 
1208: As noted in Section~\ref{s1.2}, we have the left and right analogues
1209: of the Cauchy inequality
1210: $$
1211: \abs{\Tr \ang{f,g}_R}\leq \norm{f}_R\norm{g}_R,
1212: \quad
1213: \abs{\Tr \ang{f,g}_L}\leq \norm{f}_L\, \norm{g}_L.
1214: $$
1215: Thus, $\norm{\cdot}_R$ and $\norm{\cdot}_L$ are semi-norms in $\pol$.
1216: Indeed, as noted
1217: in Section~\ref{s1.2}, they are associated to an inner product. The
1218: sets $\{ f \,\colon
1219: \norm{f}_R = 0\}$ and $\{ f \, \colon \norm{f}_L = 0\}$ are linear
1220: subspaces. Let $\pol_R$ be
1221: the completion of $\pol / \{ f \, \colon \norm{f}_R = 0\}$ (viewed as
1222: a right module over $\calM_l$)
1223: with respect to the norm $\norm{\cdot}_R$. Similarly, let $\pol_L$ be
1224: the completion of $\pol
1225: / \{ f \, \colon \norm{f}_L = 0\}$ (viewed as a left module) with
1226: respect to the norm
1227: $\norm{\cdot}_L$.
1228: 
1229: The set $\mathcal{V}$ defined in Section~\ref{s1.2} is a linear space.
1230: Let us introduce a semi-norm in $\mathcal{V}$ by
1231: \begin{equation}\label{f.vsn}
1232: \snorm{f}=\biggl\{\int d\langle f(x), \mu(x) f(x)\rangle_{\C^l}\biggr
1233: \}^{1/2}.
1234: \end{equation}
1235: Let $\mathcal{V}_0\subset \mathcal{V}$ be the linear subspace of all
1236: polynomials such that $\snorm{f}=0$ and let $\mathcal{V}_\infty$ be
1237: the completion of the quotient space $\mathcal{V}/ \mathcal{V}_0$
1238: with respect to the norm $\snorm{\cdot}$.
1239: 
1240: \begin{lemma}\label{nondeg}
1241: The following are equivalent:
1242: \begin{SL}
1243: \item[{\rm{(1)}}] $\norm{f}_R > 0$ for every non-zero $f\in \pol$.
1244: \item[{\rm{(2)}}] For all $n$, the dimension in $\pol_R$ of the set
1245: of all polynomials of degree at most $n$ is
1246: $(n+1)l^2$.
1247: \item[{\rm{(3)}}] $\norm{f}_L > 0$ for every non-zero $f\in \pol$.
1248: \item[{\rm{(4)}}] For all $n$, the dimension in $\pol_L$ of the set
1249: of all polynomials of degree at most $n$ is
1250: $(n+1)l^2$.
1251: \item[{\rm{(5)}}] For every non-zero $v \in \mathcal{V}$, we have that
1252: $\snorm{v} \not= 0$.
1253: \item[{\rm{(6)}}] For all $n$, the dimension in $\mathcal{V}_\infty$
1254: of all vector-valued polynomials
1255: of degree at most $n$ is $(n+1)l$.
1256: \end{SL}
1257: The measure $d\mu$ is called non-trivial if these equivalent
1258: conditions hold.
1259: \end{lemma}
1260: 
1261: \begin{remark} If $l=1$, these are equivalent to the usual non-
1262: triviality
1263: condition, that is, $\supp(\mu)$ is infinite. For
1264: $l>1$, we cannot define triviality in this simple way, as can be
1265: seen by looking at the direct sum of a trivial and non-trivial
1266: measure. In that case, the measure is not non-trivial in the above
1267: sense but its support is infinite.
1268: \end{remark}
1269: 
1270: \begin{proof}
1271: The equivalences  (1) $\Leftrightarrow$ (2),
1272: (3) $\Leftrightarrow$ (4), and  (5) $\Leftrightarrow$ (6)
1273: are immediate. The
1274: equivalence (1) $\Leftrightarrow$ (3) follows from \eqref{b2}.
1275: Let us prove the equivalence (1) $\Leftrightarrow$ (5).
1276: Assume that (1) holds and let $v \in\mathcal{V}$ be non-zero.
1277: Let $f\in\calM_l$ denote the matrix that has $v$ as its leftmost
1278: column and that has zero columns otherwise. Then, $0 \not= \norm{f}^2_R = \Tr \ang{f,f}_R
1279: =\snorm{v}^2$ and hence (5) holds. Now assume that (1) fails
1280: and let $f \in \pol$ be non-zero with $\norm{f}_R = 0$. Then, at
1281: least one of the column
1282: vectors of $f$ is non-zero.
1283: Suppose for simplicity that this is the first column and
1284: denote this column vector by $v$.
1285: Let $t\in\calM_l$ be the matrix $t_{ij}=\delta_{i1}\delta_{j1}$;
1286: then we have
1287: $$
1288: \norm{f}_R=0 \Rightarrow \ang{f,f}_R=0 \Rightarrow
1289: 0=\Tr( t^*\ang{f,f}_R t) =\snorm{v}^2
1290: $$
1291: and hence (5) fails.
1292: \end{proof}
1293: 
1294: 
1295: Throughout the rest of this chapter, we assume the measure $d\mu$ to
1296: be non-trivial.
1297: 
1298: \subsubsection{Monic Orthogonal Polynomials}
1299: 
1300: \begin{lemma}\label{l.b1} Let $d\mu$ be a non-trivial measure.
1301: \begin{SL}
1302: \item[{\rm{(i)}}] There exists a unique monic polynomial $P_n^R$  of
1303: degree $n$, which minimizes
1304: the norm $\norm{P_n^R}_R$.
1305: 
1306: \item[{\rm{(ii)}}] The polynomial $P_n^R$ can be equivalently defined
1307: as the monic polynomial of
1308: degree $n$ which satisfies
1309: \begin{equation}
1310: \ang{P_n^R, f}_R = {\boldsymbol 0} \quad \text{ for any } f\in \pol,
1311: \quad \deg f<n.
1312: \label{b2.2}
1313: \end{equation}
1314: 
1315: \item[{\rm{(iii)}}] There exists a unique monic polynomial $P_n^L$
1316: of degree $n$, which minimizes
1317: the norm $\norm{P_n^L}_L$.
1318: 
1319: \item[{\rm{(iv)}}] The polynomial $P_n^L$ can be equivalently defined
1320: as the monic polynomial of
1321: degree $n$ which satisfies
1322: \begin{equation}
1323: \ang{P_n^L, f}_L = {\boldsymbol 0} \quad \text{ for any } f\in \pol,
1324: \quad \deg f<n.
1325: \label{b2.3}
1326: \end{equation}
1327: 
1328: \item[{\rm{(v)}}] One has $P_n^L(x) = P_n^R(x)^\dagger$ for all  $x
1329: \in \R$ and
1330: \begin{equation}
1331: \label{b2.4} \ang{P_n^R,P_n^R}_R = \ang{P_n^L,P_n^L}_L.
1332: \end{equation}
1333: \end{SL}
1334: \end{lemma}
1335: 
1336: \begin{proof} As noted, $\pol$ has an inner product $\langle\cdot,
1337: \cdot\rangle_R$, so
1338: there is an orthogonal projection $\pi_n^{(R)}$ onto $\pol_n$
1339: discussed in Section~\ref{s1.2}. Then
1340: \begin{equation}\lb{2.6}
1341: P_n^R(x)=x^n -\pi_{n-1}^{(R)}(x^n).
1342: \end{equation}
1343: As usual, in inner product spaces, this uniquely minimizes $x^n-Q$
1344: over all $Q\in\pol_{n-1}$.
1345: It clearly obeys
1346: \begin{equation}\lb{2.6a}
1347: \Tr(\ang{P_n^R,f}_R)=0
1348: \end{equation}
1349: for all $f\in\pol_{n-1}$. But then for any matrix $\alpha$,
1350: \[
1351: \Tr(\ang{P_n^R,f}_R\alpha)=\Tr(\ang{P_n^R,f\alpha}_R)=0
1352: \]
1353: so \eqref{b2.2} holds.
1354: 
1355: This proves (i) and (ii). (iii) and (iv) are similar. To prove (v),
1356: note that
1357: $P_n^L(x)=P_n^R(x)^\dagger$ follows from the criteria \eqref{b2.2},
1358: \eqref{b2.3}.
1359: The identity \eqref{b2.4} follows from \eqref{b2}.
1360: \end{proof}
1361: 
1362: \begin{lemma}\label{l.b2}
1363: Let $\mu$ be non-trivial. For any monic polynomial $P$, we have $\det
1364: \ang{P,P}_R \not= 0$ and
1365: $\det\ang{P,P}_L \not= 0$.
1366: \end{lemma}
1367: 
1368: \begin{proof}
1369: Let $P$ be a monic polynomial of degree $n$ such that $\ang{P,P}_R$
1370: has a non-trivial
1371: kernel. Then one can find $\alpha\in\calM_l$, $\alpha\not=
1372: {\boldsymbol 0}$, such that
1373: $\alpha^\dagger \ang{P,P}_R\alpha = {\boldsymbol 0}$. It follows that
1374: $\norm{P\alpha }_R
1375: = 0$. But since $P$ is monic, the leading coefficient of $P\alpha$ is
1376: $\alpha$, so
1377: $P\alpha \not= {\boldsymbol 0}$, which contradicts the non-triviality
1378: assumption.
1379: A similar argument works for $\ang{P,P}_L$.
1380: \end{proof}
1381: 
1382: By the orthogonality of $Q_n-P_n^R$ to $P_n^R$ for any monic
1383: polynomial $Q_n$ of degree $n$, we have
1384: \begin{equation}\lb{2.7a}
1385: \ang{Q_n,Q_n}_R = \ang{Q-P_n^R, Q-P_n^R}_R + \ang{P_n^R, P_n^R}_R
1386: \end{equation}
1387: and, in particular,
1388: \begin{equation}\lb{2.7b}
1389: \ang{P_n^R,P_n^R}_R \leq \ang{Q_n,Q_n}_R
1390: \end{equation}
1391: with (by non-triviality) equality if and only if $Q_n=P_n^R$.
1392: Since $\Tr$ and $\det$ are strictly monotone on strictly positive
1393: matrices, we have the following variational principles
1394: (\eqref{2.7c} restates (i) of Lemma~\ref{l.b1}):
1395: 
1396: \begin{theorem}\lb{T2.3AA} For any monic $Q_n$ of degree $n$, we have
1397: \begin{gather}
1398: \norm{Q_n}_R \geq \norm{P_n^R}_R, \lb{2.7c} \\
1399: \det\ang{Q_n,Q_n}_R \geq \det\ang{P_n^R,P_n^R}_R \lb{2.7d}
1400: \end{gather}
1401: with equality if and only if $P_n^R=Q_n$.
1402: \end{theorem}
1403: 
1404: 
1405: \subsubsection{Expansion} \lb{s2.1.3}
1406: 
1407: \begin{theorem}\lb{T2.3A} Let $d\mu$ be non-trivial.
1408: \begin{SL}
1409: \item[{\rm{(i)}}] We have
1410: \begin{equation}\lb{2.6A}
1411: \ang{P_k^R,P_n^R}_R = \gamma_n \delta_{kn}
1412: \end{equation}
1413: for some positive invertible matrices $\gamma_n$.
1414: 
1415: \item[{\rm{(ii)}}] $\{P_k^R\}_{k=0}^n$ are a right-module basis for $
1416: \pol_n$; indeed, any $f\in\pol_n$ has
1417: a unique expansion,
1418: \begin{equation}\lb{2.6B}
1419: f=\sum_{j=0}^n P_j^R f_j^R.
1420: \end{equation}
1421: Indeed, essentially by \eqref{1.33c},
1422: \begin{equation}\lb{2.6C}
1423: f_j^R = \gamma_j^{-1} \ang{P_j^R,f}_R.
1424: \end{equation}
1425: \end{SL}
1426: \end{theorem}
1427: 
1428: 
1429: \begin{remark} There are similar formulas for $\ang{\cdot,\cdot}_L$.
1430: By \eqref{b2.4},
1431: \begin{equation}\lb{2.6D}
1432: \ang{P_k^L, P_n^L}_L =\gamma_n \delta_{kn}
1433: \end{equation}
1434: (same $\gamma_n$, which is why we use $\gamma_n$ and not $\gamma_n^R$).
1435: \end{remark}
1436: 
1437: \begin{proof}  (i) \eqref{2.6A} for $n<k$ is immediate from \eqref
1438: {b2.3} and for $n>k$ by symmetry.
1439: $\gamma_n\geq 0$ follows from \eqref{1.25a}. By Lemma~\ref{l.b2}, $
1440: \det(\gamma_n)\neq 0$, so
1441: $\gamma_n$ is invertible.
1442: 
1443: \smallskip
1444: (ii) Map $(\calM_l)^{n+1}$ to $\pol_n$ by
1445: \[
1446: \langle \alpha_0,\dots,\alpha_n\rangle \mapsto \sum_{j=0}^n P_j^R
1447: \alpha_j
1448: \equiv X (\alpha_0, \dots, \alpha_n).
1449: \]
1450: By \eqref{2.6A},
1451: \[
1452: \alpha_j=\gamma_j^{-1} \ang{P_j^R,X(\alpha_0,\dots,\alpha_n)}
1453: \]
1454: so that map is one-one. By dimension counting, it is onto.
1455: \end{proof}
1456: 
1457: \subsubsection{Recurrence Relations for Monic Orthogonal Polynomials}
1458: Denote by $\zeta_n^R$ (resp.\ $\zeta_n^L$) the coefficient of
1459: $x^{n-1}$ in $P_n^R(x)$ (resp.\ $P_n^L(x)$), that is,
1460: \begin{align*}
1461: P_n^R(x)&=x^n \bdone +\zeta_n^R x^{n-1}+\text{lower order terms},
1462: \\
1463: P_n^L(x)&=x^n \bdone +\zeta_n^L x^{n-1}+\text{lower order terms}.
1464: \end{align*}
1465: Since $P_n^R(x)^\dagger=P_n^L(x)$, we have $(\zeta_n^R)^\dagger=
1466: \zeta_n^L$.
1467: Using the parameters $\gamma_n$ of \eqref{2.6A} and $\zeta_n^R$, $
1468: \zeta_n^L$ one can write down
1469: recurrence relations for $P_n^R(x)$, $P_n^L(x)$.
1470: 
1471: \begin{lemma}
1472: \begin{SL}
1473: \item[{\rm{(i)}}] We have a commutation relation
1474: \begin{equation}
1475: \gamma_{n-1}(\zeta_n^R-\zeta_{n-1}^R)
1476: =
1477: (\zeta_n^L-\zeta_{n-1}^L)\gamma_{n-1}.
1478: \label{b.2.7.1}
1479: \end{equation}
1480: 
1481: \item[{\rm{(ii)}}] We have the recurrence relations
1482: \begin{align}
1483: xP_n^R(x)&=P_{n+1}^R(x)+P_n^R(x)(\zeta_n^R-\zeta_{n+1}^R)+P_{n-1}(x)
1484: \gamma_{n-1}^{-1}\gamma_n,
1485: \label{b.2.7.2}
1486: \\
1487: xP_n^L(x)&=P_{n+1}^L(x)+(\zeta_n^L-\zeta_{n+1}^L)P_n^L(x)+\gamma_n
1488: \gamma_{n-1}^{-1}P_{n-1}^L(x).
1489: \label{b.2.7.3}
1490: \end{align}
1491: \end{SL}
1492: \end{lemma}
1493: 
1494: 
1495: \begin{proof}
1496: (i) We have
1497: $$
1498: P_n^R(x)-xP_{n-1}^R(x)=(\zeta_n^R-\zeta_{n-1}^R)x^{n-1}+\text{lower
1499: order terms}
1500: $$
1501: and so
1502: \begin{align*}
1503: (\zeta_n^L-\zeta_{n-1}^L)\gamma_{n-1}
1504: &= (\zeta_n^R-\zeta_{n-1}^R)^\dagger \ang{P_{n-1}^R, P_{n-1}^R}_R \\
1505: & = (\zeta_n^R-\zeta_{n-1}^R)^\dagger \ang{x^{n-1}, P_{n-1}^R}_R \\
1506: & = \ang{P_n^R-xP_{n-1}^R,P_{n-1}^R}_R \\
1507: & = \ang{P_n^R,P_{n-1}^R}_R - \ang{xP_{n-1}^R,P_{n-1}^R}_R \\
1508: & = -\ang{xP_{n-1}^R,P_{n-1}^R}_R \\
1509: & = -\ang{P_{n-1}^R,xP_{n-1}^R}_R \\
1510: & = \ang{P_{n-1}^R,P_n^R-xP_{n-1}^R}_R \\
1511: & = \ang{P_{n-1}^R,x^{n-1}(\zeta_n^R-\zeta_{n-1}^R)}_R \\
1512: & = \ang{P_{n-1}^R,x^{n-1}}_R (\zeta_n^R-\zeta_{n-1}^R) \\
1513: & = \gamma_{n-1} (\zeta_n^R-\zeta_{n-1}^R).
1514: \end{align*}
1515: 
1516: \smallskip
1517: (ii)
1518: By Theorem~\ref{T2.3A},
1519: $$
1520: xP_n^R (x)=P_{n+1}^R(x)C_{n+1}+P_n^R(x)C_n+P_{n-1}^R(x)C_{n-1}+\cdots
1521: +P_0^R C_0
1522: $$
1523: with some matrices $C_0, \dots, C_{n+1}$. It is straightforward
1524: that $C_{n+1}={\boldsymbol 1}$ and $C_n=\zeta_n^R-\zeta_{n+1}^R$. By
1525: the orthogonality property \eqref{b2.2}, we find
1526: $C_0=\dots=C_{n-2} = {\boldsymbol 0}$. Finally, it is easy to
1527: calculate $C_{n-1}$:
1528: \begin{align*}
1529: \gamma_n & = \ang{P_n^R,xP_{n-1}^R}_R = \ang{xP_n^R,P_{n-1}^R}_R \\
1530: & = \ang{P_{n+1}^R+P_n^R(\zeta_n^R-\zeta_{n+1}^R)+P_{n-1}^RC_{n-1},P_
1531: {n-1}^R}_R \\
1532: & = C_{n-1}^\dagger \gamma_{n-1}
1533: \end{align*}
1534: and so, taking adjoints and using self-adjointness of $\gamma_j$, $C_
1535: {n-1} =
1536: \gamma_{n-1}^{-1}\gamma_n$. This proves \eqref{b.2.7.2}; the other
1537: relation \eqref{b.2.7.3}
1538: is obtained by conjugation.
1539: \end{proof}
1540: 
1541: 
1542: \subsubsection{Normalized Orthogonal Polynomials} We call  $p_n^R\in
1543: \pol$ a right orthonormal
1544: polynomial if $\deg p_n^R\leq n$ and
1545: \begin{gather}
1546: \ang{p_n^R,f}_R= \bdzero \text{ for every $f\in \pol$ with } \deg
1547: f<n, \label{b2.8}
1548: \\
1549: \ang{p_n^R,p_n^R}_R={\boldsymbol 1}.
1550: \label{b2.9}
1551: \end{gather}
1552: Similarly, we call  $p_n^L\in \pol$ a left orthonormal polynomial if $
1553: \deg
1554: p_n^L\leq n$ and
1555: \begin{gather}
1556: \ang{p_n^L,f}_L=\bdzero \text{ for every $f\in \pol$ with }\deg f<n,
1557: \label{b2.10}
1558: \\
1559: \ang{p_n^L,p_n^L}_L={\boldsymbol 1}.
1560: \label{b2.11}
1561: \end{gather}
1562: 
1563: \begin{lemma}\label{l.b3}
1564: Any orthonormal polynomial has the form
1565: \begin{equation}
1566: p_n^R(x)=P_n^R(x)\ang{P_n^R,P_n^R}_R^{-1/2}\sigma_n,
1567: \qquad
1568: p_n^L(x)=\tau_n\ang{P_n^L,P_n^L}_L^{-1/2}P_n^L(x)
1569: \label{b2.11.1}
1570: \end{equation}
1571: where $\sigma_n,\tau_n\in\calM_l$ are unitaries. In particular,
1572: $\deg p_n^R=\deg p_n^L=n$.
1573: \end{lemma}
1574: 
1575: \begin{proof}
1576: Let $K_n$ be the coefficient of $x^n$ in $p_n^R$.
1577: Consider the polynomial $q(x)=P_n^R(x) K_n-p_n^R(x)$,
1578: where $P_n^R$ is the monic orthogonal polynomial from Lemma~\ref{l.b1}.
1579: Then $\deg q<n$ and so from \eqref{b2.2} and \eqref{b2.8}, it follows
1580: that
1581: $\ang{q,q}_R=0$ and so $q(x)$
1582: vanishes identically. Thus, we have
1583: \begin{equation}
1584: {\boldsymbol 1}=\ang{p_n^R,p_n^R}_R
1585: =
1586: K_n^\dagger \ang{P_n^R,P_n^R}_R K_n
1587: \label{b.2.12}
1588: \end{equation}
1589: and so $\det(K_n)\not=0$.
1590: {}From \eqref{b.2.12} we get
1591: $(K_n^\dagger)^{-1}K_n^{-1}=\ang{P_n^R,P_n^R}_R$,
1592: and so $K_n K_n^\dagger=\ang{P_n^R,P_n^R}_R^{-1}$.
1593: {}From here we get $K_n=\ang{P_n^R,P_n^R}_R^{-1/2}\sigma_n$
1594: with a unitary $\sigma_n$.
1595: The proof for $p_n^L$ is similar.
1596: \end{proof}
1597: 
1598: By Theorem~\ref{T2.3A}, the polynomials $p_n^R$ form a right
1599: orthonormal module basis in
1600: $\pol_R$. Thus, for any $f\in \pol_R$, we have
1601: \begin{equation}
1602: f(x)=\sum_{m=0}^\infty p_m^R f_m,
1603: \qquad f_m=\ang{p_m^R, f}_R
1604: \label{b5}
1605: \end{equation}
1606: and the Parseval identity
1607: \begin{equation}
1608: \sum_{m=0}^\infty \Tr(f_m f_m^\dagger)=\norm{f}^2_R
1609: \label{b6}
1610: \end{equation}
1611: holds true.
1612: Obviously, since $f$ is a polynomial, there are only finitely many
1613: non-zero
1614: terms in \eqref{b5} and \eqref{b6}.
1615: 
1616: 
1617: \subsection{Block Jacobi Matrices} \lb{s2.2}
1618: 
1619: The study of block Jacobi matrices goes back at least to Krein \cite
1620: {Krein3}.
1621: 
1622: \subsubsection{Block Jacobi Matrices as Matrix Representations}
1623: 
1624: Suppose that a sequence of unitary matrices ${\boldsymbol 1} =
1625: \sigma_0, \sigma_1,
1626: \sigma_2, \ldots$ is fixed, and $p_n^R$ are defined according to
1627: \eqref{b2.11.1}. As
1628: noted above, $p_n^R$ form a right orthonormal basis in $\pol_R$.
1629: 
1630: The map $f(x)\mapsto xf(x)$ can be considered as a right homomorphism
1631: in $\pol_R$.
1632: Consider the matrix $J_{nm}$ of this homomorphism with respect to the
1633: basis $p_n^R$, that
1634: is,
1635: \begin{equation}\lb{2.22a}
1636: J_{nm}=\ang{p_{n-1}^R,xp_{m-1}^R}_R.
1637: \end{equation}
1638: Following Killip--Simon \cite{KS} and Simon \cite{S,S2,Rice}, our
1639: Jacobi matrices are
1640: indexed with $n=1,2,\dots$ but, of course, $p_n$ has $n=0,1,2,\dots$.
1641: That is why
1642: \eqref{2.22a} has $n-1$ and $m-1$.
1643: 
1644: As in the scalar case, using the orthogonality properties of $p_n^R$,
1645: we get that $J_{nm}
1646: = {\boldsymbol 0}$ if $\abs{n-m}>1$. Denote
1647: $$
1648: B_n=J_{nn}=\ang{p_{n-1}^R,x p_{n-1}^R}_R
1649: $$
1650: and
1651: $$
1652: A_n=J_{n,n+1}=J_{n+1,n}^\dagger=\ang{p_{n-1}^R,x p_{n}^R}_R.
1653: $$
1654: Then we have
1655: \begin{equation}
1656: J=
1657: \begin{pmatrix}
1658: B_1 & A_1 & {\boldsymbol 0} & \cdots \\
1659: A_1^\dagger & B_2 & A_2 & \cdots \\
1660: {\boldsymbol 0} & A_2^\dagger & B_3 &  \cdots \\
1661: \vdots & \vdots & \vdots & \ddots
1662: \end{pmatrix}
1663: \label{b6.1}
1664: \end{equation}
1665: 
1666: Applying \eqref{b5} to $f(x)=xp_n^R(x)$, we get the recurrence relation
1667: \begin{equation}
1668: xp_n^R(x)=p_{n+1}^R(x)A^\dagger_{n+1}+p_{n}^R(x)B_{n+1}+p_{n-1}^R(x)A_
1669: {n},
1670: \quad
1671: n=1,2,\dots .
1672: \label{b7}
1673: \end{equation}
1674: If we set $p_{-1}^R(x) = {\boldsymbol 0}$ and $A_0={\boldsymbol 1}$,
1675: the relation
1676: \eqref{b7} also holds for $n=0$. By \eqref{b2}, we can always pick
1677: $p_n^L$
1678: so that for $x$ real, $p_n^L(x) = p_n^R(x)^\dagger$, and thus for
1679: complex $z$,
1680: \begin{equation} \lb{2.30a}
1681: p_n^L(z) = p_n^R(\bar z)^\dagger
1682: \end{equation}
1683: by analytic continuation. By conjugating \eqref{b7}, we get
1684: \begin{equation}
1685: xp_n^L(x)=
1686: A_{n+1} p_{n+1}^L(x) + B_{n+1} p_{n}^L(x) +A_{n}^\dagger p_{n-1}^L(x),
1687: \quad
1688: n=0,1,2,\dots .
1689: \label{b8}
1690: \end{equation}
1691: Comparing this with the recurrence relations \eqref{b.2.7.2}, \eqref
1692: {b.2.7.3},
1693: we get
1694: \begin{equation}
1695: A_n=\sigma_{n-1}^\dagger\gamma_{n-1}^{-1/2}\gamma_n^{1/2}\sigma_n,
1696: \qquad
1697: B_n=\sigma_{n-1}^\dagger\gamma_{n-1}^{1/2}(\zeta_{n-1}^R-\zeta_n^R)
1698: \gamma_{n-1}^{-1/2}\sigma_{n-1}.
1699: \label{b8.00}
1700: \end{equation}
1701: In particular, $\det A_n\not=0$ for all $n$.
1702: 
1703: Notice that since $\sigma_n$ is unitary, $\abs{\det(\sigma_n)}=1$, so
1704: \eqref{b8.00} implies $\det(\gamma_n^{1/2}) = \det(\gamma_{n-1}^{1/2})
1705: \abs{\det(A_n)}$ which, by induction, implies that
1706: \begin{equation} \lb{2.32a}
1707: \det \ang{P_n^R,P_n^R} = \abs{\det(A_1\cdots A_n)}^2.
1708: \end{equation}
1709: 
1710: Any block matrix of the form \eqref{b6.1} with $B_n = B_n^\dagger$
1711: and $\det A_n \not= 0$
1712: for all $n$ will be called a block Jacobi matrix corresponding to the
1713: Jacobi parameters
1714: $A_n$ and $B_n$.
1715: 
1716: 
1717: \subsubsection{Basic Properties of Block Jacobi Matrices}
1718: 
1719: Suppose we are given a block Jacobi matrix $J$ corresponding to
1720: Jacobi parameters $A_n$
1721: and $B_n$, where $B_n = B_n^\dagger$ and $\det A_n \not= 0$ for each
1722: $n$.
1723: 
1724: Consider the Hilbert space $\h_v = \ell^2(\Z_+, \C^l)$ (here
1725: $\bbZ_+ = \{1,2,3,\dots\}$) with inner product
1726: $$
1727: \langle f,g \rangle_{\h_v} = \sum_{n=1}^\infty \langle f_n , g_n
1728: \rangle_{\C^l}
1729: $$
1730: and orthonormal basis $\{ e_{k,j} \}_{k \in \Z_+, 1 \le j \le l}$, where
1731: $$
1732: (e_{k,j})_n = \delta_{k,n} v_j
1733: $$
1734: and $\{ v_j \}_{1 \le j \le l}$ is the standard basis of $\C^l$. $J$
1735: acts on $\h_v$ via
1736: \begin{equation}\label{jacobiop}
1737: (J f)_n = A_{n-1}^\dagger f_{n-1} + B_n f_n + A_n f_{n+1}, \quad f\in
1738: \h_v
1739: \end{equation}
1740: (with $f_0 = 0$) and defines a symmetric operator on this space. Note
1741: that using
1742: invertibility of the $A_n$'s, induction shows
1743: \begin{equation}\label{cyclicity}
1744: \Span \{ e_{k,j}\, \colon 1 \le k \le n, \, 1 \le j \le l \} = \Span \{
1745: J^{k-1} e_{1,j}\, \colon 1 \le k \le n, \, 1 \le j \le l \}
1746: \end{equation}
1747: for every $n \ge 1$. We want to emphasize that elements of $\calH_v$
1748: and $\calH$ are
1749: vector-valued and matrix-valued, respectively. For this reason, we
1750: will be interested
1751: in both matrix- and vector-valued solutions of the basic difference
1752: equations.
1753: 
1754: We will consider only bounded block Jacobi matrices, that is, those
1755: corresponding to
1756: Jacobi parameters satisfying
1757: \begin{equation}\label{boundedbjm}
1758: \sup_{n} \, \Tr (A_n^\dagger A_n+B_n^\dagger B_n) < \infty.
1759: \end{equation}
1760: Equivalently,
1761: \begin{equation} \lb{2.29a}
1762: \sup_n (\norm{A_n} + \norm{B_n}) <\infty.
1763: \end{equation}
1764: In this case, $J$ is a bounded self-adjoint operator. This is
1765: equivalent to $\mu$ having
1766: compact support.
1767: 
1768: We call two Jacobi matrices $J$ and $\tilde J$ equivalent if there
1769: exists a sequence of
1770: unitaries $u_n\in\calM_l$, $n\geq 1$, with $u_1={\boldsymbol 1}$ such
1771: that $\tilde
1772: J_{nm}=u_n^\dagger J_{nm} u_m$.
1773: {}From Lemma~\ref{l.b3} it is clear that if $p_n^R$, $\tilde p_n^R$
1774: are two sequences of normalized orthogonal polynomials, corresponding
1775: to the same
1776: measure (but having different normalization), then the Jacobi matrices
1777: $J_{nm}=\ang{p_{n-1}^R, xp_{m-1}^R}_R$ and $\tilde J_{nm}=\ang{\tilde
1778: p_{n-1}^R,
1779: x\tilde p_{m-1}^R}_R$ are equivalent ($u_n=\sigma_{n-1}^\dagger \ti
1780: \sigma_{n-1}$).
1781: Thus,
1782: \begin{equation} \lb{2.35a}
1783: \ti B_n = u_n^\dagger B_n u_n, \qquad \ti A_n = u_n^\dagger A_n u_{n+1}.
1784: \end{equation}
1785: 
1786: Therefore, we have a map
1787: \begin{equation}
1788: \begin{split}
1789: \Phi \,\colon  \mu\mapsto &\{ J \,\colon J_{mn}=\ang{p_{n-1}^R,x p_
1790: {m-1}^R}_R, \, p_n^R \\
1791: &\quad \text{ correspond to $d\mu$ for some normalization}\} \label{Phi}
1792: \end{split}
1793: \end{equation}
1794: from the set of all Hermitian positive semi-definite non-trivial
1795: compactly supported measures to the set of all equivalence classes
1796: of bounded block Jacobi matrices. Below, we will see how to invert
1797: this map.
1798: 
1799: 
1800: \subsubsection{Special Representatives of the Equivalence Classes}
1801: 
1802: Let $J$ be a block Jacobi matrix with the Jacobi parameters $A_n$,
1803: $B_n$. We say that $J$
1804: is:
1805: \begin{itemize}
1806: \item
1807: of type~1, if $A_n>\bdzero$ for all $n$;
1808: \item
1809: of type~2, if $A_1 A_2 \cdots A_n> \bdzero$ for all $n$;
1810: \item
1811: of type~3, if $A_n\in {\mathcal L}$ for all $n$.
1812: \end{itemize}
1813: Here, $\mathcal L$ is the class of all lower triangular matrices
1814: with strictly positive elements on the diagonal. Type~3 is of
1815: interest because they correspond precisely to bounded Hermitian
1816: matrices with $2l+1$ non-vanishing diagonals with the extreme
1817: diagonals strictly positive; see Section~\ref{s1.4}(c). Type~2 is
1818: the case where the leading coefficients of $p_n^R$ are strictly
1819: positive definite.
1820: 
1821: \begin{theorem} \label{l.b4}
1822: \begin{SL}
1823: \item[{\rm{(i)}}] Each equivalence class of block Jacobi matrices
1824: contains exactly one element
1825: each of type~1, type~2, or type~3.
1826: 
1827: \item[{\rm{(ii)}}] Let $J$ be a block Jacobi matrix corresponding to
1828: a sequence of polynomials
1829: $p_n^R$ as in \eqref{b2.11.1}. Then $J$ is of type~2 if and only if $
1830: \sigma_n =
1831: {\boldsymbol 1}$ for all $n$.
1832: \end{SL}
1833: \end{theorem}
1834: 
1835: \begin{proof}
1836: The proof is based on the following two well-known facts:
1837: \begin{SL}
1838: \item[(a)] For any $t\in\calM_l$ with $\det(t)\not=0$, there
1839: exists a unique unitary $u\in \calM_l$ such that $tu$ is Hermitian
1840: positive semi-definite: $tu\geq 0$. \item[(b)] For any
1841: $t\in\calM_l$ with $\det(t)\not=0$, there exists a unique unitary
1842: $u\in\calM_l$ such that $tu\in \calL$.
1843: \end{SL}
1844: 
1845: We first prove that every equivalence class of block Jacobi matrices
1846: contains at least
1847: one element of type~1.
1848: For a given sequence $A_n$, let us construct a sequence $u_1=
1849: \bdone,u_2,u_3,\dots$
1850: of unitaries such that $u_n^\dagger A_n u_{n+1}\geq\bdzero$.
1851: By the existence part of (a), we find $u_2$ such that $A_1 u_2\geq
1852: \bdzero$,
1853: then find $u_3$ such that $u_2^\dagger A_2 u_3\geq\bdzero$, etc.
1854: This, together with \eqref{2.35a}, proves the statement.
1855: In order to  prove the uniqueness part, suppose we have $A_n\geq\bdzero$
1856: and $u_n^\dagger A_n u_{n+1}\geq\bdzero$ for all $n$.
1857: Then, by the uniqueness part of (a), $A_1\geq\bdzero$ and $A_1 u_2
1858: \geq \bdzero$
1859: imply $u_2=\bdone$; next, $A_2\geq\bdzero$ and $u_2^\dagger A_2
1860: u_3=A_2 u_3\geq\bdzero$
1861: imply $u_3=\bdone$, etc.
1862: 
1863: The statement (i) concerning type~$3$ can be proven in the same way,
1864: using (b) instead of
1865: (a).
1866: 
1867: The statement (i) concerning type~$2$ can be proven similarly.
1868: Existence: find $u_2$
1869: such that $A_1 u_2\geq \bdzero$, then $u_3$ such that
1870: $(A_1 u_2)(u_2^\dagger A_2 u_3)=A_1A_2u_3\geq \bdzero$, etc.
1871: Uniqueness: if
1872: $A_1\dots A_n\geq\bdzero$  and $A_1\cdots A_n u_{n+1}\geq\bdzero$,
1873: then $u_{n+1}=\bdone$.
1874: 
1875: By \eqref{b8.00}, we have $A_1 A_2 \cdots A_n = \gamma_n^{1/2}\sigma_n
1876: $ and the statement
1877: (ii) follows from the positivity of $\gamma_n$.
1878: \end{proof}
1879: 
1880: We say that a block Jacobi matrix $J$ belongs to the Nevai class if
1881: $$
1882: B_n\to \bdzero \text{ and } A_n^\dagger A_n\to {\boldsymbol 1} \text
1883: { as }n\to\infty.
1884: $$
1885: It is clear that $J$ is in the Nevai class if and only if all
1886: equivalent Jacobi matrices
1887: belong to the Nevai class.
1888: 
1889: \begin{theorem}\label{l.b5}
1890: If $J$ belongs to the Nevai class and is of type~1 or type~3, then
1891: $A_n\to 1$ as $n\to
1892: \infty$.
1893: \end{theorem}
1894: 
1895: \begin{proof}
1896: If $J$ is of type~$1$, then $A_n^\dagger A_n=A_n^2\to {\boldsymbol 1}
1897: $ clearly
1898: implies $A_n\to {\boldsymbol 1}$ since square root is continuous on
1899: positive Hermitian
1900: matrices.
1901: 
1902: Suppose $J$ is of type~$3$. We shall prove that $A_n\to \bdone$ by
1903: considering the rows
1904: of the matrix $A_n$ one by one, starting from the $l$th row. Denote $
1905: (A_n)_{jk} =
1906: a_{j,k}^{(n)}$. We have
1907: $$
1908: (A_n^\dagger A_n)_{ll}=(a_{l,l}^{(n)})^2\to1,
1909: \text{ and so } a_{l,l}^{(n)} \to1.
1910: $$
1911: Then, for any $k<l$, we have
1912: $$
1913: (A_n^\dagger A_n)_{lk}=a^{(n)}_{l,l}a^{(n)}_{l,k}\to 0,
1914: \text{ and so } a_{l,k}^{(n)} \to 0.
1915: $$
1916: Next, consider the $(l-1)$st row. We have
1917: $$
1918: (A_n^\dagger A_n)_{l-1,l-1}=(a_{l-1,l-1}^{(n)})^2+\abs{a_{l,l-1}^
1919: {(n)}}^2 \to 1
1920: $$
1921: and so, using the previous step, $a_{l-1,l-1}^{(n)}\to1$ as $n\to
1922: \infty$.
1923: Then for all $k<l-1$, we have
1924: $$
1925: (A_n^\dagger A_n)_{l-1,k}=\overline{a_{l-1,l-1}^{(n)}}\, a^{(n)}_{l-1,k}
1926: +\overline{a_{l,l-1}^{(n)}} \, a^{(n)}_{l,k}\to 0
1927: $$
1928: and so, using the previous steps, $a_{l-1,k}\to 0$.
1929: Continuing this way, we get $a^{(n)}_{j,k}\to \delta_{j,k}$ as required.
1930: \end{proof}
1931: 
1932: It is an interesting open question if this result also applies to the
1933: type~2 case.
1934: 
1935: \subsubsection{Favard's Theorem}
1936: 
1937: Here we construct an inverse of the mapping $\Phi$ (defined by \eqref
1938: {Phi}).
1939: Thus, $\Phi$ sets up a bijection between non-trivial measures of compact
1940: support and equivalence classes of bounded block Jacobi matrices.
1941: 
1942: Before proceeding to do this, let us prove:
1943: 
1944: \begin{lemma}\label{l.b6}
1945: The mapping $\Phi$ is injective.
1946: \end{lemma}
1947: 
1948: \begin{proof}
1949: Let $\mu$ and $\tilde \mu$ be two Hermitian positive semi-definite
1950: non-trivial compactly supported measures. Suppose that
1951: $\Phi(\mu)=\Phi(\tilde \mu)$.
1952: 
1953: Let $p_n^R$ and $\tilde p_n^R$ be normalized orthogonal polynomials
1954: corresponding to $\mu$ and $\tilde\mu$. Suppose that the normalization
1955: both for $p_n^R$ and for $\tilde p_n^R$ has been chosen such that
1956: $\sigma_n={\boldsymbol 1}$ (see \eqref{b2.11.1}), that is, type~2.
1957: {}From Lemma~\ref{l.b4} and the assumption $\Phi(\mu)=\Phi(\tilde \mu)
1958: $ it
1959: follows that the corresponding Jacobi matrices coincide, that is,
1960: $\ang{p_n^R,xp_m^R}_R=\ang{\tilde p_n^R,x\tilde p_m^R}_R$
1961: for all $n$ and $m$.
1962: Together with the recurrence relation \eqref{b7} this yields $p_n^R=
1963: \tilde p_n^R$
1964: for all $n$.
1965: 
1966: For any $n \geq 0$, we can represent $x^n$ as
1967: $$
1968: x^n = \sum_{k=0}^{n} p_k^R(x) C^{(n)}_k =\sum_{k=0}^{n} \tilde p^R_k
1969: (x) \tilde C^{(n)}_k.
1970: $$
1971: The coefficients $C^{(n)}_k$ and $\tilde C^{(n)}_k$ are completely
1972: determined by the
1973: coefficients of the polynomials $p_n^R$ and $\tilde p_n^R$ and so
1974: $C^{(n)}_k=\tilde C^{(n)}_k$ for all $n$ and $k$.
1975: 
1976: For the moments of the measure $\mu$, we have
1977: $$
1978: \int x^n d\mu(x)= \ang{{\boldsymbol 1}, x^n}_R
1979: =
1980: \sum_{k=0}^n \ang{ {\boldsymbol 1}, p^R_k C_k^{(n)}}_R
1981: =
1982: \ang{ {\boldsymbol 1}, {\boldsymbol 1}}_R \, C_0^{(n)} =C_0^{(n)}.
1983: $$
1984: Since the same calculation is valid for the measure $\tilde \mu$, we get
1985: $$
1986: \int x^n d\mu(x)=\int x^n d\tilde \mu(x)
1987: $$
1988: for all $n$. It follows that
1989: $$
1990: \int f(x) d\mu(x)g(x)=\int f(x) d\tilde \mu(x) g(x)
1991: $$
1992: for all matrix-valued polynomials $f$ and $g$, and so the measures $
1993: \mu$ and $\tilde \mu$
1994: coincide.
1995: \end{proof}
1996: 
1997: We can now construct the inverse of the map $\Phi$. Let a block
1998: Jacobi matrix $J$ be given. By a version of the spectral theorem
1999: for self-adjoint operators with finite multiplicity (see, e.g.,
2000: \cite[Sect.~72]{AG}), there exists a matrix-valued measure $d\mu$
2001: with
2002: \begin{equation}
2003: \langle e_{1,j} , f(J) e_{1,k} \rangle_{\h_v} = \int f(x) \, d\mu_
2004: {j,k}(x)
2005: \label{defmu}
2006: \end{equation}
2007: and an isometry
2008: $$
2009: R \, \colon \h_v \to L^2(\R,d\mu;\C^l)
2010: $$
2011: such that (recall that $\{v_j\}$ is the standard basis in $\C^l$)
2012: \begin{equation}\label{constant}
2013: [R e_{1,j}](x) = v_j, \quad 1 \le j \le l,
2014: \end{equation}
2015: and, for any $g\in \h_v$, we have
2016: \begin{equation}\label{diagonal}
2017: (RJg)(x) = x (Rg)(x).
2018: \end{equation}
2019: If the Jacobi matrices $J$ and $\tilde J$ are equivalent, then we
2020: have $\tilde J = U^*\! J
2021: U$ for some $U = \oplus_{n=1}^\infty u_n$, $u_1={\boldsymbol 1}$. Thus,
2022: $$
2023: \langle e_{1,j} , f(\tilde J) e_{1,k} \rangle_{\h_v} = \langle U e_
2024: {1,j} , f(J) Ue_{1,k}
2025: \rangle_{\h_v} = \langle  e_{1,j} , f(J) e_{1,k} \rangle_{\h_v}
2026: $$
2027: and so the measures corresponding to $J$ and $\tilde J$ coincide.
2028: Thus, we have a map
2029: \begin{equation}
2030: \Psi \colon \{\tilde J \colon \tilde J \text{ is equivalent to } J\}
2031: \mapsto \mu
2032: \label{Psi}
2033: \end{equation}
2034: from the set of all equivalence classes of bounded block Jacobi
2035: matrices to the set of all Hermitian positive semi-definite
2036: compactly supported measures.
2037: 
2038: \begin{theorem}\label{favard}
2039: \begin{SL}
2040: \item[{\rm{(i)}}] All measures in the image of the map $\Psi$ are non-
2041: degenerate.
2042: \item[{\rm{(ii)}}] $\Phi\circ \Psi=\text{id}$.
2043: \item[{\rm{(iii)}}] $\Psi\circ \Phi=\text{id}$.
2044: \end{SL}
2045: \end{theorem}
2046: 
2047: \begin{proof}
2048: (i) To put things in the right context, we first recall that
2049: $\norm{\cdot}_{\h_v}$ is a norm (rather than a semi-norm), whereas
2050: $\snorm{\cdot}$  on $\mathcal{V}$ (cf.~\eqref{f.vsn}) is, in
2051: general, a semi-norm. Using the assumption that $\det(A_k)\not=0$
2052: for all $k$ (which is included in our definition of a Jacobi
2053: matrix), we will prove that $\snorm{\cdot}$ is in fact a norm.
2054: More precisely, we will prove that  $\snorm{p}>0$ for any
2055: non-zero polynomial $p\in\mathcal{V}$; by Lemma~\ref{nondeg} this will
2056: imply that $\mu$ is non-degenerate.
2057: 
2058: Let $p\in\mathcal{V}$ be a non-zero polynomial, $\deg p=n$. Notice
2059: that \eqref{constant} and \eqref{diagonal} give
2060: \begin{equation}\label{conjugation}
2061: [R J^k e_{1,j}](x) = x^k v_j
2062: \end{equation}
2063: for every $k \ge 0$ and $1 \le j \le l$. This shows that $p$ can
2064: be represented as $p=Rg$, where $g=\sum_{k=0}^n J^k f_k$, and
2065: $f_0, \dots, f_n$ are vectors in $\calH_v$ such that $\langle f_i,
2066: e_{j,k}\rangle_{\calH_v}=0$ for all $i=0,\dots,n$, $j\geq2$,
2067: $k=1,\dots, l$ (i.e., the only non-zero components of $f_j$ are in
2068: the first $\bbC^l$ in $\calH_v$). Assumption $\deg p=n$ means
2069: $f_n\not=0$.
2070: 
2071: Since $R$ is isometric, we have $\snorm{p}=\norm{g}_{\h_v}$, and
2072: so we have to prove that $g\not=0$.  Indeed, suppose that $g=0$.
2073: Using the assumption $\det(A_k)\not=0$ and the tri-diagonal nature
2074: of $J$, we  see that $\sum_{k=0}^n J^k f_k=0$ yields $f_n=0$,
2075: contrary to our assumption.
2076: 
2077: \smallskip
2078: (ii) Consider the elements $R e_{n,k}\in L^2(\R,d\mu;\C^l)$. First
2079: note that, by
2080: \eqref{cyclicity} and \eqref{conjugation}, $R e_{n,k}$ is a
2081: polynomial of degree at most
2082: $n-1$. Next, by the unitarity of $R$, we have
2083: \begin{equation}
2084: \langle R e_{n,k}, R e_{m,j}\rangle_{L^2(\R,d\mu;\C^l)} = \delta_
2085: {m,n} \delta_{k,j}.
2086: \label{ortho}
2087: \end{equation}
2088: Let us construct matrix-valued polynomials $q_n(x)$, using $R e_{n,
2089: 1}, R e_{n,2}, \dots,
2090: R e_{n,l}$ as columns of $q_{n-1}(x)$:
2091: $$
2092: [q_{n-1}(x)]_{j,k} =  [R e_{n,k}(x)]_j.
2093: $$
2094: We have $\deg q_n \leq n$ and $\ang{q_m,q_n}_R = \delta_{m,n}
2095: {\boldsymbol 1}$; the last
2096: relation is just a reformulation of \eqref{ortho}. Hence the $q_n$'s
2097: are right
2098: normalized orthogonal polynomials with respect to the measure $d\mu$.
2099: We find
2100: \begin{align*}
2101: J_{nm} & = [ \langle e_{n,j}, J e_{m,k} \rangle_{\h_v} ]_{1 \le j,k
2102: \le l} \\
2103: & = [ \langle R e_{n,j}, R J e_{m,k} \rangle_{L^2(\R,d\mu;\C^l)} ]_{1
2104: \le j,k \le l} \\
2105: & = [ \langle R e_{n,j}, x R e_{m,k} \rangle_{L^2(\R,d\mu;\C^l)} ]_{1
2106: \le j,k \le l} \\
2107: & =  [ \langle [q_{n-1}(x)]_{\bddot,j}, x
2108: [q_{m-1}(x)]_{\bddot,k} \rangle_{L^2(\R,d\mu;\C^l)}]_{1 \le j,k \le
2109: l} \\
2110: & =  \ang{q_{n-1} , x q_{m-1}}_R
2111: \end{align*}
2112: as required.
2113: 
2114: \smallskip
2115: (iii)
2116: Follows from (ii) and from Lemma~\ref{l.b6}.
2117: \end{proof}
2118: 
2119: 
2120: 
2121: \subsection{The $m$-Function}
2122: 
2123: \subsubsection{The Definition of the $m$-Function}
2124: 
2125: We denote the Borel transform of $d\mu$ by $m$:
2126: \begin{equation}
2127: m(z) = \int \frac{d\mu(x)}{x-z}\,  ,\qquad \Ima z > 0. \label{b.8.0}
2128: \end{equation}
2129: It is a matrix-valued Herglotz function, that is, it is analytic
2130: and obeys $\Ima m(z) > 0$. For information on matrix-valued
2131: Herglotz functions, see \cite {GT} and references therein.
2132: Extensions to operator-valued Herglotz functions can be found in
2133: \cite{GKMT}.
2134: 
2135: \begin{lemma}
2136: Suppose $d\mu$ is given, $p_n^R$ are right normalized orthogonal
2137: polynomials, and $J$ is
2138: the associated block Jacobi matrix. Then,
2139: \begin{equation}\label{mfitop}
2140: m(z) = \ang{p_0^R,(x-z)^{-1}p_0^R}_R
2141: \end{equation}
2142: and
2143: \begin{equation}\label{mfitoj}
2144: m(z) = \langle e_{1,\bddot}, (J-z)^{-1} e_{1,\bddot} \rangle_{\h_v}.
2145: \end{equation}
2146: \end{lemma}
2147: 
2148: \begin{proof}
2149: Since $p_0^R = {\boldsymbol 1}$, \eqref{mfitop} is just a way of
2150: rewriting the definition
2151: of $m$. The second identity, \eqref{mfitoj}, is a consequence of
2152: \eqref{defmu} and
2153: Theorem~\ref{favard}(iii).
2154: \end{proof}
2155: 
2156: 
2157: \subsubsection{Coefficient Stripping}
2158: 
2159: If $J$ is a block Jacobi matrix corresponding to invertible $A_n$'s
2160: and Hermitian
2161: $B_n$'s, we denote the $k$-times stripped block Jacobi matrix,
2162: corresponding to $\{
2163: A_{k+n} , B_{k+n} \}_{n \ge 1}$, by $J^{(k)}$. That is,
2164: $$
2165: J^{(k)} =
2166: \begin{pmatrix}
2167: B_{k+1} & A_{k+1} & {\boldsymbol 0} & \cdots \\
2168: A_{k+1}^\dagger & B_{k+2} & A_{k+2} & \cdots \\
2169: {\boldsymbol 0} & A_{k+2}^\dagger & B_{k+3} &  \cdots \\
2170: \vdots & \vdots & \vdots & \ddots
2171: \end{pmatrix}.
2172: $$
2173: The $m$-function corresponding to $J^{(k)}$ will be denoted by $m^
2174: {(k)}$. Note that, in
2175: particular, $J^{(0)} = J$ and $m^{(0)} = m$.
2176: 
2177: \begin{proposition}
2178: Let $J$ be a block Jacobi matrix with $\sigma_\mathrm{ess}(J)
2179: \subseteq [a,b]$. Then, for
2180: every $\varepsilon > 0$, there is $k_0 \ge 0$ such that for $k \ge k_0
2181: $, we have that
2182: $\sigma(J^{(k)}) \subseteq [a-\varepsilon,b+\varepsilon]$.
2183: \end{proposition}
2184: 
2185: \begin{proof}
2186: This is an immediate consequence of (the proof of) \cite[Lemma~1]{Den}.
2187: \end{proof}
2188: 
2189: \begin{proposition}[Due to Aptekarev--Nikishin \cite{AN}]\label{strip}
2190: We have that
2191: $$
2192: m^{(k)}(z)^{-1}  = B_{k+1} - z - A_{k+1} m^{(k+1)}(z) A_{k+1}^\dagger
2193: $$
2194: for $\Ima z > 0$ and $k \ge 0$.
2195: \end{proposition}
2196: 
2197: \begin{proof} It suffices to handle the case $k=0$.
2198: Given \eqref{mfitoj}, this is a special case of a general formula for
2199: $2\times 2$ block
2200: operator matrices, due to Schur \cite{Schur}, that reads
2201: $$
2202: \begin{pmatrix}
2203: A & B \\
2204: C & D
2205: \end{pmatrix}^{-1} = \begin{pmatrix}
2206: (A-BD^{-1}C)^{-1} & -A^{-1} B (D-CA^{-1}B)^{-1} \\
2207: -D^{-1} C (A-BD^{-1}C)^{-1} & (D-CA^{-1}B)^{-1}
2208: \end{pmatrix}
2209: $$
2210: and which is readily verified. Here $A=B_1-z$, $B=A_1$, $C=A_1^\dagger
2211: $, and $D=J^{(1)}-z$.
2212: \end{proof}
2213: 
2214: 
2215: \subsection{Second Kind Polynomials} Define the second kind
2216: polynomials by $q_{-1}^R(z) =
2217: -{\boldsymbol 1}$,
2218: $$
2219: q_n^R(z)=\int_\R d\mu(x) \frac{p_n^R(z)-p_n^R(x)}{z-x},
2220: \quad n=0,1,2,\dots .
2221: $$
2222: As is easy to see, for $n\geq 1$, $q_n^R$ is a polynomial of degree
2223: $n-1$. For future reference,
2224: let us display the first several polynomials $p_n^R$ and $q_n^R$:
2225: \begin{alignat}{3}
2226: p_{-1}^R(x) &= {\boldsymbol 0}, \quad & p_{0}^R(x) &= {\boldsymbol
2227: 1}, \quad & p_{1}^R(x) &=
2228: (x-B_1)A_1^{-1}, \label{b10} \\
2229: q_{-1}^R(x) &= -{\boldsymbol 1}, \quad & q_{0}^R(x) &= {\boldsymbol
2230: 0}, \quad & q_{1}^R(x) &=
2231: A_1^{-1}. \label{b11}
2232: \end{alignat}
2233: The polynomials $q_n^R$ satisfy the equation (same form as \eqref{b7})
2234: \begin{equation}
2235: xq_n^R(x)=q_{n+1}^R(x)A_{n+1}^\dagger+q_{n}^R(x)B_{n+1}+q_{n-1}^R(x)A_
2236: {n},
2237: \quad
2238: n=0,1,2,\dots .
2239: \label{b12}
2240: \end{equation}
2241: For $n=0$, this can be checked by a direct substitution of
2242: \eqref{b11}. For $n\geq1$, as in the scalar case, this can be
2243: checked by taking \eqref{b7} for $x$ and for $z$, subtracting,
2244: dividing by $x-z$, integrating over $d\mu$, and taking into
2245: account the orthogonality relation
2246: $$
2247: \int d\mu(x) p_n^R(x) = {\boldsymbol 0}, \quad n\geq1.
2248: $$
2249: Finally, let us define
2250: $$
2251: \psi_n^R(z)=q_n^R(z)+m(z)p_n^R(z).
2252: $$
2253: According to the definition of $q_n^R$, we have
2254: $$
2255: \psi_n^R(z)=\ang{f_z, p_n^R}_R,
2256: \quad
2257: f_z(x)=(x-\bar{z})^{-1}.
2258: $$
2259: By the Parseval identity, this shows that for all $\Ima z>0$, the
2260: sequence $\psi_n^R(z)$ is in $\ell^2$, that is,
2261: \begin{equation}\lb{2.49ax}
2262: \sum_{n=0}^\infty \Tr (\psi_n^R(z)^\dagger \psi_n^R(z)) <\infty .
2263: \end{equation}
2264: 
2265: In the same way, we define $q_{-1}^L(z)=-{\boldsymbol 1}$,
2266: $$
2267: q_n^L(z)=\int_\R  \frac{p_n^L(z)-p_n^L(x)}{z-x}d\mu(x)\, ,
2268: \quad n=0,1,2,\dots
2269: $$
2270: and $\psi_n^L(z)=q_n^L(z)+p_n^L(z)m(z)$.
2271: 
2272: 
2273: \subsection{Solutions to the Difference Equations} For $\Ima z>0$,
2274: consider the solutions
2275: to the equations
2276: \begin{align}
2277: z u_n(z)&=
2278: \sum_{m=1}^\infty u_m(z) J_{mn},
2279: \quad
2280: n=2,3,\dots,
2281: \label{b15}
2282: \\
2283: z v_n(z)&=
2284: \sum_{m=1}^\infty  J_{nm} v_m(z) ,
2285: \quad
2286: n=2,3,\dots .
2287: \label{b16}
2288: \end{align}
2289: Clearly, $u_n(z)$ solves \eqref{b15} if and only if $v_n(z)=(u_n(\bar
2290: z))^\dagger$
2291: solves \eqref{b16}. In the above, we normally intend $z$ as a fixed
2292: parameter
2293: but it then can be used as a variable. That is, $z$ is fixed and $u_n
2294: (z)$ is a fixed sequence, not a $z$-dependent function. A statement
2295: like $v_n
2296: (z)=
2297: (u_n(\bar z))^\dagger$ means if $u_n$ is a sequence solving \eqref
2298: {b15} for
2299: $z=\bar z_0$, then $v_n$ obeys \eqref{b16} for $z=z_0$. Of course, if
2300: $u_n(z)$
2301: is a function of $z$ in a region, we can apply our estimates to all $z
2302: $ in the
2303: region. For any solution $\{u_n(z)\}_{n=1}^\infty$ of \eqref{b15},
2304: let us define
2305: \begin{equation}\lb{2.49a}
2306: u_0(z)=zu_1(z)-u_1(z) B_1-u_2(z)A_1^\dagger.
2307: \end{equation}
2308: With this definition, the equation \eqref{b15} for $n=1$ is
2309: equivalent to
2310: $u_0(z)=0$. In the same way, we set
2311: $$
2312: v_0(z)=zv_1(z)-B_1 v_1(z) -A_1 v_2(z).
2313: $$
2314: 
2315: 
2316: \begin{lemma}
2317: Let $\Ima z>0$ and suppose $\{u_n(z)\}_{n=0}^\infty$ solves
2318: \eqref{b15} {\rm{(}}for $n\geq 2${\rm{)}} and \eqref{2.49a} and
2319: belongs to $\ell^2$.
2320: Then
2321: \begin{equation}
2322: (\Ima z) \sum_{n=1}^\infty \Tr (u_n(z)^\dagger u_n(z))
2323: =
2324: -\Ima \Tr (u_1(z) u_0(z)^\dagger).
2325: \label{b17}
2326: \end{equation}
2327: In particular, $u_n(z) = \alpha p_{n-1}^R(z)$ is in $\ell^2$ only if
2328: $\alpha=\bdzero$.
2329: \end{lemma}
2330: 
2331: \begin{proof}
2332: Denote $s_n=\Tr (u_n(z) A_{n-1}^\dagger u_{n-1}(z)^\dagger)$. Here
2333: $A_0=\bdone$. Multiplying \eqref{b15} for $n\geq 2$ and
2334: \eqref{2.49a} for $n=1$ by $u_n(z)^\dagger$ on the right, taking
2335: traces, and summing over $n$, we get
2336: $$
2337: z\sum_{n=1}^N \Tr(u_n(z) u_n(z)^\dagger)
2338: =
2339: \sum_{n=1}^N s_{n+1}
2340: + \sum_{n=1}^N \Tr(u_n(z) B_{n+1} u_n(z)^\dagger)
2341: + \sum_{n=1}^N \overline{s_n}\, .
2342: $$
2343: Taking imaginary parts and letting $N\to \infty$, we obtain
2344: \eqref{b17} since the middle sum is real and the outer sums cancel
2345: up to boundary terms. Applying \eqref{b17} to $u_n(z)=\alpha p_{n-1}^R
2346: (z)$,
2347: we get zero in the right-hand side:
2348: $$
2349: (\Ima z)   \sum_{n=1}^\infty \Tr (\alpha p_{n-1}^R(z) p_{n-1}^R(z)^
2350: \dagger\alpha^\dagger)
2351: =0
2352: $$
2353: and hence $\alpha=\bdzero$ since $p_{0}^R=\bdone$.
2354: \end{proof}
2355: 
2356: 
2357: \begin{theorem}\label{t.b1}
2358: Let $\Ima z>0$.
2359: \begin{SL}
2360: \item[{\rm{(i)}}] Any  solution $\{u_n(z)\}_{n=0}^\infty$ of \eqref
2361: {b15} {\rm{(}}for $n\geq 2${\rm{)}}
2362: can be represented as
2363: \begin{equation}
2364: u_n(z)=a p_{n-1}^R(z) + b q_{n-1}^R(z)
2365: \label{b18}
2366: \end{equation}
2367: for suitable $a,b\in\calM_l$. In fact, $a=u_1(z)$ and $b=-u_0(z)$.
2368: 
2369: \item[{\rm{(ii)}}] A sequence  \eqref{b18} satisfies \eqref{b15} for
2370: all $n\geq 1$ if and only if $b=0$.
2371: 
2372: \item[{\rm{(iii)}}] A sequence \eqref{b18} belongs to $\ell^2$ if and
2373: only if $u_n(z)=c\psi_{n-1}^R(z)$ for some
2374: $c\in\calM_l$. Equivalently, a sequence \eqref{b18} belongs to $\ell^2
2375: $ if and only if $u_1(z)+u_0(z)m(z)=0$.
2376: \end{SL}
2377: \end{theorem}
2378: 
2379: \begin{proof}
2380: (i) Let $u_n(z)$ be a solution to \eqref{b15}. Consider
2381: $$
2382: \tilde u_n(z)
2383: =
2384: u_n(z) - u_1(z) p_{n-1}^R(z) + u_0(z)q_{n-1}^R(z).
2385: $$
2386: Then $\tilde u_n(z)$ also solves \eqref{b15} and
2387: $\tilde u_0(z)=\tilde u_1(z)=0$.
2388: It follows that $\tilde u_n(z)=0$ for all $n$. This proves (i).
2389: 
2390: \smallskip
2391: (ii) A direct substitution of \eqref{b18} into \eqref{b15} for $n=1$
2392: yields the statement.
2393: 
2394: \smallskip
2395: (iii)
2396: We already know that $c\psi_{n-1}^R$ is an $\ell^2$ solution.
2397: Suppose that $u_n(z)$ is an $\ell^2$ solution to \eqref{b15}. Rewrite
2398: \eqref{b18} as
2399: $$
2400: u_n(z)=(a-bm(z))p_{n-1}^R(z)+b \psi_{n-1}^R(z).
2401: $$
2402: Since $\psi_n^R$ is in $\ell^2$ and $cp_n^R$ is not in $\ell^2$,
2403: we get $a=bm(z)$, which is equivalent to $u_1(z)+u_0(z)m(z)=0$ or
2404: to $u_n(z)=b\psi_{n-1}^R(z)$.
2405: \end{proof}
2406: 
2407: By conjugation, we obtain:
2408: 
2409: \begin{theorem}\label{t.b2}
2410: Let $\Ima z>0$.
2411: \begin{SL}
2412: \item[{\rm{(i)}}] Any  solution $\{v_n(z)\}_{n=0}^\infty$ of \eqref
2413: {b16} {\rm{(}}for $n\geq 2${\rm{)}}
2414: can be represented as
2415: \begin{equation}
2416: v_n(z)= p_{n-1}^L(z)a + q_{n-1}^L(z)b .
2417: \label{b19}
2418: \end{equation}
2419: In fact, $a=v_1(z)$ and $b=-v_0(z)$.
2420: 
2421: \item[{\rm{(ii)}}] A sequence  \eqref{b19} satisfies \eqref{b16} for
2422: all $n\geq1$ if and only if $b=0$.
2423: 
2424: \item[{\rm{(iii)}}] A sequence \eqref{b19} belongs to $\ell^2$ if
2425: and only if $v_n(z)=\psi_{n-1}^L(z)c$ for some $c\in\calM_l$.
2426: Equivalently, a sequence \eqref{b19} belongs to $\ell^2$ if and
2427: only if $v_1(z)+m(z) v_0(z)=0$.
2428: \end{SL}
2429: \end{theorem}
2430: 
2431: 
2432: \subsection{Wronskians and the Christoffel--Darboux Formula}
2433: 
2434: For any two $\calM_l$-valued sequences $u_n$, $v_n$, define the
2435: Wronskian by
2436: \begin{equation}
2437: W_n(u,v)=u_{n} A_n v_{n+1} - u_{n+1} A_n^\dagger v_{n}. \label{b8.1}
2438: \end{equation}
2439: 
2440: Note that $W_{n}(u, v)= - W_{n}(v^\dagger, u^\dagger)^\dagger$. If
2441: $u_n(z)$ and $v_n(z)$
2442: are solutions to \eqref{b15} and \eqref{b16}, then by a direct
2443: calculation, we see that
2444: $W_n(u(z),v(z))$ is independent of $n$. Put differently, if both $u_n
2445: (z)$ and $v_n(z)$
2446: are solutions to \eqref{b15},
2447: then  $W_n(u(z),v(\bar z)^\dagger)$ is independent of $n$.
2448: Or, if both $u_n(z)$ and $v_n(z)$ are solutions to \eqref{b16},
2449: then  $W_n(u(\bar z)^\dagger, v(z))$ is independent of $n$.
2450: In particular, by a direct evaluation for $n=0$, we get
2451: \begin{align*}
2452: W_n(p_{\bddot\, -1}^R(z), p_{\bddot\, -1}^R(\bar{z})^\dagger)&= W_n(q_
2453: {\bddot\, -1}^R(z),
2454: q_{\bddot\, -1}^R(\bar{z})^\dagger)= \bdzero, \\
2455: W_n(p_{\bddot\, -1}^L(\bar{z})^\dagger, p_{\bddot \, -1}^L(z))&= W_n
2456: (q_{\bddot\, -1}^L(\bar{z})^\dagger,
2457: q_{\bddot\, -1}^L(z))= \bdzero, \\
2458: W_n(p_{\bddot\, -1}^R(z), q_{\bddot\, -1}^R(\bar{z})^\dagger)&= W_n(p_
2459: {\bddot\, -1}^L(\bar{z})^\dagger,
2460: q_{\bddot\, -1}^L(z))={\boldsymbol 1}.
2461: \end{align*}
2462: Let both $u(z)$ and $v(z)$ be solutions to \eqref{b15} of the type
2463: \eqref{b18}, namely,
2464: $$
2465: u_n(z)=a p_{n-1}^R(z) + b q_{n-1}^R(z), \quad v_n(z)=c p_{n-1}^R(z) +
2466: d q_{n-1}^R(z).
2467: $$
2468: Then the above calculation implies
2469: $$
2470: W_n(u(z), v(\bar{z})^\dagger)=ad^\dagger - b c^\dagger.
2471: $$
2472: 
2473: \begin{theorem}[CD Formula]\lb{T2.18}
2474: For any $x,y\in\C$ and $n\geq1$, one has
2475: \begin{equation}
2476: (x-y)\sum_{k=0}^n p_k^R(x) p_k^L(y) = -W_{n+1}(p_{\bddot\, -1}^R(x),
2477: p_{\bddot\, -1}^L(y)). \label{b8.2}
2478: \end{equation}
2479: \end{theorem}
2480: \begin{proof}
2481: Multiplying \eqref{b7} by $p_n^L(y)$ on the right and \eqref{b8}
2482: (with $y$ in place of $x$) by $p_n^R(x)$ on the left and
2483: subtracting, we get
2484: $$
2485: (x-y) p_n^R(x) p_n^R(y) = W_n(p_{\bddot\, -1}^R(x), p_{\bddot\, -1}^L
2486: (y))
2487: -W_{n+1}(p_{\bddot\, -1}^R(x), p_{\bddot\, -1}^L(y)).
2488: $$
2489: Summing over $n$ and noting that $W_0(p^R(x), p^L(y))=0$, we get the
2490: required statement.
2491: \end{proof}
2492: 
2493: 
2494: \subsection{The CD Kernel}\lb{2.7}
2495: 
2496: The CD kernel is defined for $z,w\in\bbC$ by
2497: \begin{align}
2498: K_n(z,w) &= \sum_{k=0}^n p_k^R(z) p_k^R(\bar w)^\dagger \lb{2.54a} \\
2499: &= \sum_{k=0}^n p_k^L (\bar z)^\dagger p_k^L(w). \lb{2.54b}
2500: \end{align}
2501: 
2502: \eqref{2.54b} follows from \eqref{2.54a} and \eqref{2.30a}.
2503: Notice that $K$ is independent of the choices $\sigma_n,\tau_n$ in
2504: \eqref{b2.11.1} and that \eqref{b8.2} can be written
2505: \begin{equation} \lb{2.54c}
2506: (z-\bar w)K_n(z,w) = -W_{n+1} (p_{\bddot\, -1}^R(z), p_{\bddot\, -1}^R
2507: (\bar w)^\dagger).
2508: \end{equation}
2509: 
2510: The independence of $K_n$ of $\sigma,\tau$ can be understood by
2511: noting that if $f_m$ is given by \eqref{b5}, then
2512: \begin{equation} \lb{2.54cA}
2513: \int K_n(z,w)\, d\mu(w) f(w) =\sum_{m=0}^n p_m^R (z) f_m
2514: \end{equation}
2515: so $K$ is the kernel of the orthogonal projection onto polynomials
2516: of degree up to $n$, and so intrinsic. Similarly, if
2517: $f_m^{(L)}=\ang{f, p_m^L}_L$, so
2518: \begin{equation} \lb{2.54d}
2519: f(x)=\sum_{m=0}^\infty f_m^{(L)} p_m^L(x),
2520: \end{equation}
2521: then, by \eqref{2.54b},
2522: \begin{equation} \lb{2.54e}
2523: \int f(z)\, d\mu(z) K_n(z,w) =\sum_{m=0}^n f_m^{(L)} p_m^L (w).
2524: \end{equation}
2525: 
2526: One has
2527: \begin{equation} \lb{2.54f}
2528: \int K_n(z,w)\, d\mu(w) K_n(w,\zeta) =K_n (z,\zeta)
2529: \end{equation}
2530: as can be checked directly and which is an expression of the fact
2531: that the map in \eqref{2.54cA} is a projection, and so its own
2532: square.
2533: 
2534: We will let $\pi_n$ be the map of $L^2 (d\mu)$ to itself given by
2535: \eqref{2.54cA} (or by \eqref{2.54e}). \eqref{2.54c} can then be
2536: viewed (for $z,w\in\bbR$) as an expression of the integral kernel
2537: of the commutator $[J,\pi_n]$, which leads to another proof of it
2538: \cite{Sim-cd}.
2539: 
2540: Let $J_{n;F}$ be the finite $nl\times nl$ matrix obtained from $J$
2541: by taking the top leftmost $n2$ blocks. It is the matrix of
2542: $\pi_{n-1} M_x\pi_{n-1}$ where $M_x$ is multiplication by $x$ in
2543: the $\{p_j^R\}_{j=0}^{n-1}$ basis. For $y\in\bbC$ and
2544: $\gamma\in\bbC^l$, let $\varphi_{n,\gamma}(y)$ be the vector whose
2545: components are
2546: \begin{equation} \lb{2.54g}
2547: (\varphi_{n,\gamma}(y))_j =p_{j-1}^L(y)\gamma
2548: \end{equation}
2549: for $j=1,2,\dots, n$. We claim that
2550: \begin{equation} \lb{2.54h}
2551: [(J_{n;F}-y)\varphi_{n,\gamma}(y)]_j = -\delta_{jn} A_{n} p_n^L(y)\gamma
2552: \end{equation}
2553: as follows immediately from \eqref{b8}.
2554: 
2555: This is intimately related to \eqref{b8.2} and \eqref{2.54c}. For
2556: recalling $J$ is the matrix in
2557: $p^R$ basis, $\varphi_{n,\gamma}(y)$ corresponds to the function
2558: \[
2559: \sum_{j=0}^{n-1} p_j^R(x) (\varphi_{n,\gamma}(y))_{j-1} = K_n (x,y)
2560: \gamma .
2561: \]
2562: As we will see in the next two sections, \eqref{2.54h} has important
2563: consequences.
2564: 
2565: \subsection{Christoffel Variational Principle} \lb{s2.7A}
2566: 
2567: There is a matrix version of the Christoffel variational principle
2568: (see Nevai \cite{NevFr} for a
2569: discussion of uses in the scalar case; this matrix case is discussed
2570: by Duran--Polo \cite{DP}):
2571: 
2572: \begin{theorem}\lb{T2.18A} For any non-trivial $l\times l$ matrix-
2573: valued measure, $d\mu$, on
2574: $\bbR$, we have that for any $n$, any $x_0\in\bbR$, and matrix
2575: polynomials $Q_n(x)$ of degree
2576: at most $n$ with
2577: \begin{equation}\lb{2.66a}
2578: Q_n(x_0)=\bdone,
2579: \end{equation}
2580: we have that
2581: \begin{equation}\lb{2.66b}
2582: \ang{Q_n, Q_n}_R \geq K_n(x_0,x_0)^{-1}
2583: \end{equation}
2584: with equality if and only if
2585: \begin{equation}\lb{2.66c}
2586: Q_n(x) =K_n(x,x_0) K_n(x_0,x_0)^{-1}.
2587: \end{equation}
2588: \end{theorem}
2589: 
2590: \begin{remark} \eqref{2.66b} also holds for $\ang{\cdot,\cdot}_L$ but
2591: the minimizer is then
2592: $K_n(x_0,x_0)^{-1} K_n(x,x_0)$.
2593: \end{remark}
2594: 
2595: \begin{proof} Let $Q_n^{(0)}$ denote the right-hand side of \eqref
2596: {2.66c}. Then for any
2597: polynomial $R_n$ of degree at most $n$, we have
2598: \begin{equation}\lb{2.66d}
2599: \ang{Q_n^{(0)}, R_n}_R = K_n(x_0,x_0)^{-1} R_n(x_0)
2600: \end{equation}
2601: because of \eqref{2.54cA}. Since $Q_n(x_0)=Q_n^{(0)}(x_0)=\bdone$, we
2602: conclude
2603: \begin{equation}\lb{2.66e}
2604: \ang{Q_n-Q_n^{(0)},Q_n-Q_n^{(0)}}_R =\ang{Q_n,Q_n}_R - K_n(x_0,x_0)^{-1}
2605: \end{equation}
2606: from which \eqref{2.66b} is immediate and, given the supposed non-
2607: triviality, the uniqueness of
2608: minimizer.
2609: \end{proof}
2610: 
2611: With this, one easily gets an extension of a result of M\'at\'e--Nevai \cite{MN} to MOPRL.
2612: (They had it for scalar OPUC. For OPUC, it is in M\'at\'e--Nevai--Totik \cite{MNT88}
2613: on $[-1,1]$ and in Totik \cite{Tot} for general OPRL. The result
2614: below can be proven using
2615: polynomial mappings \`a la Totik \cite{Tot-acta} or Jost solutions
2616: \`a la Simon \cite{Weak-cd}.)
2617: 
2618: \begin{theorem} \lb{T2.18B} Let $d\mu$ be a non-trivial $l\times l$
2619: matrix-valued measure on
2620: $\bbR$ with compact support, $E$. Let $I=(a,b)$ be an open interval
2621: with $I\subset E$. Then for
2622: Lebesgue a.e.\ $x\in I$,
2623: \begin{equation}\lb{2.66f}
2624: \limsup (n+1) K_n(x,x)^{-1} \leq w(x).
2625: \end{equation}
2626: \end{theorem}
2627: 
2628: \begin{remark} This is intended in terms of expectations in any fixed
2629: vector.
2630: \end{remark}
2631: 
2632: We state this explicitly since we will need it in Section~\ref
2633: {s5.3A}, but we note that
2634: the more detailed results of M\'at\'e--Nevai--Totik \cite{MNT88},
2635: Lubinsky \cite{Lub},
2636: Simon \cite{2ext}, and Totik \cite{Tot-prep} also extend.
2637: 
2638: 
2639: 
2640: \subsection{Zeros} \lb{s2.8}
2641: 
2642: We next look at zeros of $\det(P_n^L(z))$, which we will prove soon
2643: is also $\det(P_n^R(z))$. Following
2644: \cite{DL,Sinap}, we will identify these zeros with eigenvalues of $J_
2645: {n;F}$. It is not obvious a priori
2646: that these zeros are real and, unlike the scalar situation, where the
2647: classical arguments on the
2648: zeros rely on orthogonality, we do not know how to get reality just
2649: from that (but see the remark
2650: after Theorem~\ref{T2.18E}).
2651: 
2652: \begin{lemma}\lb{L2.18A} Let $C(z)$ be an $l\times l$ matrix-valued
2653: function analytic near $z=0$. Let
2654: \begin{equation} \lb{2.54i}
2655: k=\dim(\ker(C(0))).
2656: \end{equation}
2657: Then $\det(C(z))$ has a zero at $z = 0$ of order at least $k$.
2658: \end{lemma}
2659: 
2660: \begin{remarks} 1. Even in the $1\times 1$ case, where $k=1$, the
2661: zeros can clearly be of higher
2662: order than $k$ since $c_{11}(z)$ can have a zero of any order!
2663: 
2664: \smallskip
2665: 2. The temptation to take products of eigenvalues will lead at best
2666: to a complicated proof
2667: as the cases $C(z)=\left( \begin{smallmatrix} 0&z\\ 1&0 \end
2668: {smallmatrix}\right)$ and $C(z)
2669: =\left(\begin{smallmatrix} 0&z^2 \\ 1&0 \end{smallmatrix}\right)$
2670: illustrate.
2671: \end{remarks}
2672: 
2673: \begin{proof} Let $e_1, \dots , e_l$ be an orthonormal basis with $e_1,
2674: \dots , e_k\in \ker(C(0))$.
2675: By Hadamard's inequality (see Bhatia \cite{Bha}),
2676: \begin{align*}
2677: \abs{\det(C(z))} &\leq \norm{C(z) e_1} \cdots \norm{C(z) e_l} \\
2678: &\leq C\abs{z}^k
2679: \end{align*}
2680: since $\norm{C(z)e_j}\leq C\abs{z}$ if $j=1, \dots, k$ and $\norm{C
2681: (z) e_j}\leq d$ for $j=k+1, \dots, l$.
2682: \end{proof}
2683: 
2684: The following goes back at least to \cite{Dean-Martin}; see also
2685: \cite{DL,SB07,Sinap,SinapVanAsche}.
2686: \begin{theorem}\lb{T2.18BB} We have that
2687: \begin{equation} \lb{2.54j}
2688: \det_{\bbC^l}(P_n^L(z)) =\det_{\bbC^{nl}}(z-J_{n;F}).
2689: \end{equation}
2690: \end{theorem}
2691: 
2692: \begin{proof} By \eqref{2.54h}, if $\gamma$ is such that $p_n^L(y)
2693: \gamma =0$, then $\varphi_{n,\gamma}(y)$ is an eigenvector for
2694: $J_{n;F}$ with eigenvalue $y$. Conversely, if $ \varphi$ is an
2695: eigenvector and $\gamma$ is defined as that vector in $\bbC^l$
2696: whose components are the first $l$ components of $\varphi$, then a
2697: simple inductive argument shows $\varphi=\varphi_{n,\gamma}(y)$
2698: and then, by \eqref{2.54h} and the fact that $A_{n}$ is
2699: invertible, we see that $p_n^L(y)\gamma=0$. This shows that for
2700: any $y$,
2701: \begin{equation} \lb{2.54k}
2702: \dim\ker (P_n^L(y))=\dim\ker (J_{n;F}-y).
2703: \end{equation}
2704: 
2705: By Lemma~\ref{L2.18A}, if $y$ is any eigenvalue of $J_{n;F}$ of
2706: multiplicity $k$, then $\det(P_n^L(z))$ has a zero of order at
2707: least $k$ at $y$. Now let us consider the polynomials in $z$ on
2708: the left and right in \eqref{2.54j}. Since $J_{n;F}$ is Hermitian,
2709: the sum of the multiplicities of the zeros on the right is $nl$.
2710: Since the polynomial on the left is of degree $nl$, by a counting
2711: argument it has the same set of zeros with the same multiplicities
2712: as the polynomial on the right. Since both polynomials are monic,
2713: they are equal.
2714: \end{proof}
2715: 
2716: \begin{corollary}\lb{C2.18C} All the zeros of $\det(P_n^L(z))$ are
2717: real. Moreover,
2718: \begin{equation} \lb{2.54L}
2719: \det(P_n^R(z))=\det(P_n^L(z)).
2720: \end{equation}
2721: \end{corollary}
2722: 
2723: \begin{proof} Since $J_{n;F}$ is Hermitian, all its eigenvalues are
2724: real, so \eqref{2.54j} implies
2725: the zeros of $\det(P_n^L(z))$ are real. Thus, since the polynomial is
2726: monic,
2727: \begin{equation} \lb{2.54m}
2728: \ol{\det(P_n^L(\bar z))} =\det(P_n^L(z)).
2729: \end{equation}
2730: By Lemma~\ref{l.b1}(v), we have
2731: \begin{equation} \lb{2.54n}
2732: P_n^R(z)=P_n^L(\bar z)^\dagger
2733: \end{equation}
2734: since both sides are analytic and agree if $z$ is real. Thus,
2735: \[
2736: \det(P_n^R(z)) =\det(P_n^L(\bar z)^\dagger) =\ol{\det(P_n^L(\bar z))}
2737: \]
2738: proving \eqref{2.54L}.
2739: \end{proof}
2740: 
2741: 
2742: The following appeared before in \cite{SinapVanAsche}; see also
2743: \cite {SB07}.
2744: \begin{corollary} \lb{C2.18D} Let $\{x_{n,j}\}_{j=1}^{nl}$ be the
2745: zeros of $\det(P_n^L(x))$ counting
2746: multiplicity ordered by
2747: \begin{equation} \lb{2.54o}
2748: x_{n,1} \leq x_{n,2} \leq \cdots\leq x_{n,nl}.
2749: \end{equation}
2750: Then
2751: \begin{equation} \lb{2.54p}
2752: x_{n+1,j}\leq x_{n,j} \leq x_{n+1,j+l}.
2753: \end{equation}
2754: \end{corollary}
2755: 
2756: \begin{remarks} 1. This is interlacing if $l=1$ and replaces it for
2757: general $l$.
2758: 
2759: \smallskip
2760: 2. Using $A_n$ invertible, one can show the inequalities in \eqref
2761: {2.54p} are strict.
2762: \end{remarks}
2763: 
2764: \begin{proof} The min-max principle \cite{RS4} says that
2765: \begin{equation} \lb{2.84a}
2766: x_{n,j} = \max_{\substack{ L\subset\bbC^{nl} \\ \dim(L)\leq j-1}}\,
2767: \min_{\substack{f\in L^\perp \\ \norm{f}=1}}
2768: \jap{f,J_{n;F}f}_{\bbC^{nl}}.
2769: \end{equation}
2770: If $P\colon \bbC^{(n+1)l} \to \bbC^{nl}$ is the natural projection, then
2771: $\jap{Pf,J_{n+1;F}Pf}_{\bbC^{(n+1)l}} = \jap{Pf,J_{n;F} Pf}_{\bbC^{nl}}$
2772: and as $L$ runs through all subspaces of $\bbC^{(n+1)l}$ dimension at most
2773: $j-1$, $P[L]$ runs through all subspaces of dimension at most $j-1$ in $\bbC^{nl}$,
2774: so \eqref{2.84a} implies $x_{n+1,j}\leq x_{n,j}$.
2775: Using the same argument on $-J_{n;F}$ and
2776: $-J_{n+1;F}$ shows $x_j(-J_{n;F}) \geq x_j (-J_{n+1;F})$. But $x_j (-
2777: J_{n;F})=-x_{nl+1-j}
2778: (J_{n;F})$ and $x_j(-J_{n+1;F})=-x_{(n+1)l+1-j} (J_{n+1;F})$. That
2779: yields the other inequality.
2780: \end{proof}
2781: 
2782: \subsection{Lower Bounds on $p$ and the Stieltjes--Weyl
2783: Formula for $m$}\lb{s2.9}
2784: 
2785: Next, we want to exploit \eqref{2.54h} to prove uniform lower bounds
2786: on $\norm{p_n^L(y)\gamma}$
2787: when $y\notin \cvh(\sigma(J))$, the convex hull of the support of $J
2788: $, and thereby uniform bounds
2789: $\norm{p_n^L(y)^{-1}}$. We will then use that to prove that for $z
2790: \notin\sigma(J)$, we have
2791: \begin{equation} \lb{2.54q}
2792: m(z)=\lim_{n\to\infty} -p_n^L(z)^{-1} q_n^L(z)
2793: \end{equation}
2794: the matrix analogue of a formula that spectral theorists associate
2795: with Weyl's definition of the
2796: $m$-function \cite{Weyl}, although for the discrete case, it goes
2797: back at least to Stieltjes
2798: \cite{Stie}.
2799: 
2800: We begin by mimicking an argument from \cite{EqMC}. Let $H=\cvh(\sigma
2801: (J))=[c-D,c+D]$ with
2802: \begin{equation} \lb{2.54r}
2803: D=\tfrac12\, \diam(H).
2804: \end{equation}
2805: By the definition of $A_n$,
2806: \begin{equation} \lb{2.54s}
2807: \norm{A_n} =\norm{\ang{p_{n-1}^R, (x-c)p_n^R}_R}\leq D.
2808: \end{equation}
2809: Suppose $y\notin H$ and let
2810: \begin{equation} \lb{2.54t}
2811: d=\dist(y,H).
2812: \end{equation}
2813: By the spectral theorem, for any vector $\varphi\in\h_v$,
2814: \begin{equation} \lb{2.54u}
2815: \abs{\jap{\varphi, (J-y)\varphi}_{\h_v}}\geq d\norm{\varphi}^2.
2816: \end{equation}
2817: 
2818: By \eqref{2.54h}, with $\varphi=\varphi_{n,\gamma}(y)$,
2819: \begin{equation} \lb{2.54v}
2820: \abs{\jap{\varphi, (J-y)\varphi}} \leq \norm{A_{n}}\, \norm{p_n^L(y)
2821: \gamma}\,
2822: \norm{p_{n-1}^L(y)\gamma}
2823: \end{equation}
2824: while
2825: \begin{equation} \lb{2.54w}
2826: \norm{\varphi}^2 =\sum_{j=0}^{n-1} \, \norm{p_j^L(y)\gamma}^2.
2827: \end{equation}
2828: As in \cite[Prop.~2.2]{EqMC}, we get:
2829: 
2830: \begin{theorem}\lb{T2.18E} If $y\notin H$, for any $\gamma$,
2831: \begin{equation} \lb{2.54x}
2832: \norm{p_n^L(y)\gamma}\geq \bigl(\tfrac{d}{D}\bigr)
2833: \bigl(1+\bigl(\tfrac{d}{D}\bigr)^2\bigr)^{(n-1)/2}\norm{\gamma}.
2834: \end{equation}
2835: In particular,
2836: \begin{equation} \lb{2.54y}
2837: \norm{p_n^L(y)^{-1}} \leq \tfrac{D}{d}\, .
2838: \end{equation}
2839: \end{theorem}
2840: 
2841: \begin{remark} \eqref{2.54x} implies $\det(p_n^L(y))\neq 0$ if $\Ima
2842: y >0$, providing
2843: another proof that its zeros are real.
2844: \end{remark}
2845: 
2846: By the discussion after \eqref{b12}, if $\Ima z>0$, $q_n^L(z) + p_n^L
2847: (z)m(z)$ is in $\ell^2$,
2848: so goes to zero. Since $p_n^L(z)^{-1}$ is bounded, we obtain:
2849: 
2850: \begin{corollary}\lb{C2.18F} For any $z\in\bbC_+=\{z : \Ima
2851: z>0\}$,
2852: \begin{equation} \lb{2.54z}
2853: m(z)=\lim_{n\to\infty} -p_n^L(z)^{-1} q_n^L(z).
2854: \end{equation}
2855: \end{corollary}
2856: 
2857: \begin{remark} This holds for $z\notin H$.
2858: \end{remark}
2859: 
2860: Taking adjoints using \eqref{2.54n} and $m(z)^\dagger=m(\bar z)$, we
2861: see that
2862: \begin{equation} \lb{2.54aa}
2863: m(z)=\lim_{n\to\infty}  -q_n^R(z) p_n^R(z)^{-1}.
2864: \end{equation}
2865: 
2866: \eqref{2.54z} and \eqref{2.54aa} are due to \cite{Dur96}, which uses the
2867: proof based on the Gauss--Jacobi quadrature formula.
2868: 
2869: 
2870: \subsection{Wronskians of Vector-Valued Solutions}\lb{s2.10}
2871: 
2872: Let $\alpha,\beta$ be two vector-valued solutions $(\bbC^l)$ of
2873: \eqref {b16} for $n=2,3,\dots$. Define their scalar Wronskian as
2874: (Euclidean inner product on $\bbC^l $)
2875: \begin{equation} \lb{2.54bb}
2876: W_n(\alpha,\beta) =\jap{\alpha_{n}, A_{n}\beta_{n+1}} - \jap{A_{n}
2877: \alpha_{n+1}, \beta_{n}}
2878: \end{equation}
2879: for $n=2,3,\dots$. One can obtain two matrix solutions by using $
2880: \alpha$ or $\beta$ for one column
2881: and $0$ for the other columns. The scalar Wronskian is just a matrix
2882: element of the resulting
2883: matrix Wronskian, so $W_n$ is constant (as can also be seen by direct
2884: calculation). Here is an
2885: application:
2886: 
2887: \begin{theorem}\lb{T2.18F} Let $z_0 \in \R \setminus \sigma_\ess(J)$.
2888: For $k = 0,1$, let $q_k$ be the
2889: multiplicity of $z_0$ as an eigenvalue of $J^{(k)}$. Then, $q_0 + q_1
2890: \le l$.
2891: \end{theorem}
2892: 
2893: \begin{proof} If $\ti\beta$ is an eigenfunction for $J^{(1)}$ and we
2894: define $\beta$ by
2895: \begin{equation} \lb{2.54cc}
2896: \beta_n =\begin{cases} 0, & n=1, \\
2897: \ti\beta_{n-1}, & n\geq 2,
2898: \end{cases}
2899: \end{equation}
2900: then $\beta$ solves \eqref{b16} for $n\geq2$. If $\alpha$ is an
2901: eigenfunction for $J=J^{(0)}$, it also solves \eqref{b16}. Since
2902: $\alpha_n\to 0$, $\beta_n\to 0$, and $A_n$ is bounded,
2903: $W_n(\alpha,\beta)\to 0$ as $n\to\infty$ and so it is identically
2904: zero. But since $\beta_1 =0 $,
2905: \begin{equation} \lb{2.54dd}
2906: \bdzero=W_1(\alpha,\beta) =\jap{\alpha_1, A_1\beta_2} = \jap
2907: {\alpha_1, A_1\ti\beta_1}.
2908: \end{equation}
2909: 
2910: Let $V^{(k)}$ be the set of values of eigenfunctions of $J^{(k)}$ at
2911: $n=1$. \eqref{2.54dd}
2912: says
2913: \begin{equation} \lb{2.54ee}
2914: V^{(0)}\subset [A_1 V^{(1)}]^\perp .
2915: \end{equation}
2916: Since $q_k=\dim(V^{(k)})$ and $A_1$ is non-singular, \eqref{2.54ee}
2917: implies that $q_0\leq l-q_1$.
2918: \end{proof}
2919: 
2920: \subsection{The Order of Zeros/Poles of $m(z)$} \lb{s2.11}
2921: 
2922: \begin{theorem}\lb{T2.18H} Let $z_0 \in \R \setminus \sigma_\ess(J)$.
2923: For $k = 0,1$, let $q_k$ be the
2924: multiplicity of $z_0$ as an eigenvalue of $J^{(k)}$. If $q_1 - q_0
2925: \geq0$, then $\det(m(z))$ has a
2926: zero at $z=z_0$ of order $q_1 - q_0$. If $q_1-q_0<0$, then  $\det(m
2927: (z))$ has a pole at $z=z_0$
2928: of order $q_0 - q_1$.
2929: \end{theorem}
2930: 
2931: \begin{remarks} 1. To say $\det(m(z))$ has a zero at $z=z_0$ of order
2932: $0$ means it is finite
2933: and non-vanishing at $z_0$!
2934: 
2935: \smallskip
2936: 2. Where $d\mu$ is a direct sum of scalar measures, so is $m(z)$, and
2937: $\det(m(z))$ is then a product.
2938: In the scalar case, $m(z)$ has a pole at $z_0$ if $J^{(0)}$ has $z_0$
2939: as eigenvalue and a zero at
2940: $z_0$ if $J^{(1)}$ has $z_0$ as eigenvalue. In the direct sum case,
2941: we see there can be
2942: cancellations, which helps explain why $q_1-q_0$ occurs.
2943: 
2944: \smallskip
2945: 3. Formally, one can understand this theorem as follows. Cramer's
2946: rule suggests $\det(m(z))=
2947: \det (J^{(1)}-z)/\det(J^{(0)}-z)$. Even though $\det(J^{(k)}-z)$ is
2948: not well-defined in the
2949: infinite case, we expect a cancellation of zeros of order $q_1$ and
2950: $q_0$. For $z_0\notin H$,
2951: the convex hull of $\sigma(J^{(0)})$, one can use \eqref{2.54z} to
2952: prove the theorem following
2953: this intuition. Spurious zeros in gaps of $\sigma (J^{(0)})$ make
2954: this strategy difficult in gaps.
2955: 
2956: \smallskip
2957: 4. Unlike in Lemma~\ref{L2.18A}, we can write $m$ as a product of
2958: eigenvalues and analyze that directly because $m(x)$ is
2959: self-adjoint for $x$ real, which sidesteps some of the problems
2960: associated with non-trivial Jordan normal forms.
2961: 
2962: \smallskip
2963: 5. This proof gives another demonstration that $q_0 + q_1\leq l$.
2964: \end{remarks}
2965: 
2966: \begin{proof} $m(z)$ has a simple pole at $z=z_0$ with a residue
2967: which is rank $q_0$ and
2968: strictly negative definite on its range. Let
2969: \[
2970: f(z)=(z-z_0) m(z).
2971: \]
2972: $f$ is analytic near $z_0$ and self-adjoint for $z$ real and near
2973: $z_0$. Thus, by eigenvalue perturbation theory \cite{Kato,RS4},
2974: $f(z)$ has $l$ eigenvalues $\rho_1(z), \dots, \rho_l(z)$ analytic
2975: near $z_0$ with $\rho_1, \rho_2, \dots, \rho_{q_0}$ non-zero at
2976: $z_0$ and $\rho_{q_0+1}, \dots, \rho_l$ zero at $z_0$.
2977: 
2978: Thus, $m(z)$ has $l$ eigenvalues near $z_0$,
2979: $\lambda_j(z)=\rho_j(z)/(z-z_0)$, where $\lambda_1, \dots,
2980: \lambda_{q_0}$ have simple poles and the others are regular.
2981: 
2982: By Proposition~\ref{strip}, $m(z)^{-1}$ has a simple pole at $z_0$
2983: with residue of rank $q_1$ (because $A_1$ is non-singular),
2984: so $m(z)^{-1}$ has
2985: $q_1$ eigenvalues with poles. That means $q_1$ of
2986: $\lambda_{q_0+1}, \dots, \lambda_l$ have simple zeros at $z_0$ and
2987: the others are non-zero. Thus, $\det(m(z))=\prod_{j=1}^l
2988: \lambda_j(z)$ has a pole/zero of order $q_0-q_1$.
2989: \end{proof}
2990: 
2991: 
2992: \subsection{Resolvent of the Jacobi Matrix} Consider the matrix
2993: $$
2994: G_{nm}(z)= \ang{p_{n-1}^R, (x-z)^{-1}p_{m-1}^R}_R.
2995: $$
2996: 
2997: \begin{theorem} \lb{T2.27}
2998: One has
2999: \begin{equation}
3000: G_{nm}(z)=
3001: \begin{cases}
3002: \psi_{n-1}^L(z) p_{m-1}^R(z), & \text{ if $n\geq m$},
3003: \\
3004: p_{n-1}^L(z) \psi_{m-1}^R(z), & \text{ if $n\leq m$}.
3005: \end{cases}
3006: \end{equation}
3007: \end{theorem}
3008: \begin{proof}
3009: We have
3010: \begin{alignat*}{2}
3011: \sum_{m=1}^\infty G_{km}(z) J_{mn} &= z G_{kn}(z),
3012: \quad & n &\not= k,
3013: \\
3014: \sum_{m=1}^\infty  J_{nm} G_{mk}(z) &= z G_{nk}(z),
3015: \quad &  n& \not = k.
3016: \end{alignat*}
3017: Fix $k\geq0$ and let $u_m(z)=G_{km}(z)$. Then $u_m(z)$ satisfies the
3018: equation \eqref{b15}
3019: for $n\not =k$, and so we can use Theorem~\ref{t.b1} to describe this
3020: solution.
3021: First suppose $k>1$. As $u_m$ is an $\ell^2$ solution and $u_m$
3022: satisfies \eqref{b15} for
3023: $n=1$, we have
3024: \begin{equation}
3025: G_{km}(z)=
3026: \begin{cases}
3027: a_k(z) p_{m-1}^R(z), & m\leq k,
3028: \\
3029: b_k(z) \psi_{m-1}^R(z), & m\geq k.
3030: \end{cases}
3031: \label{b20}
3032: \end{equation}
3033: If $k=1$, \eqref{b20} also holds true. For $m\geq k$, this follows by
3034: the same
3035: argument, and for $m=k=1$, this is a trivial statement.
3036: Next, similarly, let us consider $v_m(z)=G_{mk}(z)$.
3037: Then $v_m(z)$ solves \eqref{b16} and so, using Theorem~\ref{t.b2}, we
3038: obtain
3039: \begin{equation}
3040: G_{mk}(z)=
3041: \begin{cases}
3042: p_{m-1}^L(z) c_k(z), & m\leq k,
3043: \\
3044: \psi_{m-1}^L(z) d_k(z) , & m\geq k.
3045: \end{cases}
3046: \label{b21}
3047: \end{equation}
3048: Comparing \eqref{b20} and \eqref{b21}, we find
3049: \begin{align*}
3050: a_k(z) p_{m-1}^R(z) & =  \psi_{k-1}^L(z) d_m(z),
3051: \\
3052: b_k(z) \psi_{m-1}^R(z) & =  p_{k-1}^L(z) c_m(z).
3053: \end{align*}
3054: As $p_0^R=p_0^L={\boldsymbol 1}$,
3055: it follows that
3056: $$
3057: a_k(z)=\psi_{k-1}^L(z) d_1(z),
3058: \qquad
3059: c_m(z)=b_1(z) \psi_{m-1}^R(z)
3060: $$
3061: and so we obtain
3062: \begin{equation}
3063: G_{nm}(z)=
3064: \begin{cases}
3065: \psi_{n-1}^L(z)d_1(z) p_{m-1}^R(z) & \text{ if $n\geq m$},
3066: \\
3067: p_{n-1}^L(z) b_1(z) \psi_{m-1}^R(z) & \text{ if $n\leq m$}.
3068: \end{cases}
3069: \label{b22}
3070: \end{equation}
3071: It remains to prove that
3072: \begin{equation}
3073: b_1(z)=d_1(z)={\boldsymbol 1}.
3074: \label{b23}
3075: \end{equation}
3076: Consider the case $m=n=1$.
3077: By the definition of the resolvent,
3078: $$
3079: G_{11}(z)
3080: =
3081: \ang{p_0^R,(J-z)^{-1}p_0^R}_R
3082: =
3083: \int \frac{d\mu(x)}{x-z}
3084: =
3085: m(z).
3086: $$
3087: On the other hand, by \eqref{b22},
3088: \begin{align*}
3089: G_{11}(z)&=\psi_0^L(z)d_1(z) p_0^R(z)=m(z) d_1(z),
3090: \\
3091: G_{11}(z)&=p_0^L(z) b_1(z) \psi_0^R(z)=b_1(z) m(z),
3092: \end{align*}
3093: which proves \eqref{b23}.
3094: \end{proof}
3095: 
3096: 
3097: 
3098: \section{Matrix Orthogonal Polynomials on the Unit Circle} \lb{s3}
3099: 
3100: \subsection{Definition of MOPUC} \lb{s3.1}
3101: 
3102: In this chapter, $\mu$ is an $l\times l$ matrix-valued measure on $
3103: \partial\bbD$.
3104: $\ang{\cdot,\cdot}_R$ and $\ang{\cdot,\cdot}_L$ are defined as in
3105: the MOPRL case. Non-triviality is defined as for MOPRL. We will
3106: always assume $\mu$ is non-trivial.  We define monic matrix
3107: polynomials $\Phi_n^R, \Phi_n^L$ by applying Gram--Schmidt to
3108: $\{\bdone, z\bdone, \dots\}$, that is, $\Phi_n^R$ is the unique
3109: matrix polynomial $z^n \bdone +$ lower order with
3110: \begin{equation} \lb{3.1}
3111: \ang{z^k\bdone, \Phi_n^R}_R =0 \quad k=0,1,\dots, n-1.
3112: \end{equation}
3113: We will define the normalized MOPUC shortly. We will only consider
3114: the analogue of what we called type~1 for MOPRL because only those
3115: appear to be useful. Unlike in the scalar case, the monic
3116: polynomials do not appear much because it is for the normalized,
3117: but not monic, polynomials that the left and right Verblunsky
3118: coefficients are the same.
3119: 
3120: 
3121: \subsection{The Szeg\H{o} Recursion}
3122: 
3123: Szeg\H{o} \cite{Szb} included the scalar Szeg\H{o} recursion for the
3124: first time. It seems likely
3125: that Geronimus had it independently shortly after Szeg\H{o}. Not
3126: knowing of the connection
3127: with this work, Levinson \cite{Lev} rederived the recursion but with
3128: matrix coefficients!
3129: So the results of this section go back to 1947.
3130: 
3131: For a matrix polynomial $P_n$ of degree $n$, we define the reversed
3132: polynomial $P_n^*$ by
3133: \begin{equation} \lb{3.1a}
3134: P_n^*(z) = z^n P_n(1/\bar{z})^\dagger.
3135: \end{equation}
3136: Notice that
3137: \begin{equation} \lb{3.2x}
3138: (P_n^*)^* = P_n
3139: \end{equation}
3140: and for any $\alpha\in\calM_l$,
3141: \begin{equation} \lb{3.2}
3142: (\alpha P_n)^* = P_n^* \alpha^\dagger, \qquad (P_n\alpha)^* = \alpha^
3143: \dagger P_n^*.
3144: \end{equation}
3145: 
3146: \begin{lemma}
3147: We have
3148: \begin{equation} \lb{3.1ax}
3149: \ang{f , g}_L = \ang{g , f}_L^\dagger \, ,
3150: \qquad
3151: \ang{f , g}_R = \ang{g , f}_R^\dagger
3152: \end{equation}
3153: and
3154: \begin{equation} \lb{3.1b}
3155: \ang{f^* , g^*}_L = \ang{f , g}_R^\dagger \, ,
3156: \quad
3157: \ang{f ^*, g^*}_R = \ang{f ,g}_L^\dagger \, .
3158: \end{equation}
3159: \end{lemma}
3160: 
3161: \begin{proof}
3162: The first and second identities  follow immediately from the
3163: definition. The third identity is
3164: derived as follows:
3165: \begin{align*}
3166: \ang{f^* , g^*}_L & = \int e^{in\theta} g(\theta)^\dagger \,
3167: d\mu(\theta) \, (e^{in\theta} f(\theta)^\dagger)^\dagger \\
3168: & = \int e^{in\theta} g(\theta)^\dagger \, d\mu(\theta) \, e^{-in
3169: \theta} f(\theta) \\
3170: & = \int g(\theta)^\dagger \, d\mu(\theta) \, f(\theta) \\
3171: & = \ang{g , f}_R \\
3172: & = \ang{f , g}_R^\dagger .
3173: \end{align*}
3174: The proof of the last identity is analogous.
3175: \end{proof}
3176: 
3177: \begin{lemma}\label{l.c1}
3178: If $P_n$ has degree $n$ and is left-orthogonal with respect to $z
3179: \bdone, \ldots , z^n \bdone$, then
3180: $P_n = c (\Phi_n^R)^*$ for some suitable matrix $c$.
3181: \end{lemma}
3182: 
3183: \begin{proof}
3184: By assumption,
3185: $$
3186: \bdzero = \ang{P_n , z^j \bdone}_L = \ang{(z^j \bdone)^* , P_n^*}_R =
3187: \ang{z^{n-j} \bdone , P_n^*}_R \; \text{ for } 1
3188: \le j \le n.
3189: $$
3190: Thus, $P_n^*$ is right-orthogonal with respect to $\bdone, z\bdone,
3191: \ldots, z^{n-1} \bdone$ and hence it
3192: is a right-multiple of $\Phi_n^R$. Consequently, $P_n$ is a left-
3193: multiple of
3194: $(\Phi_n^R)^*$.
3195: \end{proof}
3196: 
3197: Let us define normalized orthogonal matrix polynomials by
3198: $$
3199: \varphi_0^L=\varphi_0^R=\bdone,\quad
3200: \varphi_n^L = \kappa_n^L \Phi_n^L \quad \text{ and } \quad \varphi_n^R =
3201: \Phi_n^R\kappa_n^R
3202: $$
3203: where the $\kappa$'s are defined according to the normalization
3204: condition
3205: $$
3206: \ang{\varphi_n^R,\varphi_m^R}_R=\delta_{nm}{\boldsymbol 1}
3207: \quad
3208: \ang{\varphi_n^L,\varphi_m^L}_L=\delta_{nm}{\boldsymbol 1}
3209: $$
3210: along with (a type~1 condition)
3211: \begin{equation}\lb{3.4a}
3212: \kappa_{n+1}^L (\kappa_n^L)^{-1} > \bdzero \quad \text{and} \quad
3213: (\kappa_n^R)^{-1}
3214: \kappa_{n+1}^R > \bdzero.
3215: \end{equation}
3216: Notice that $\kappa_{n}^L$ are determined by the normalization
3217: condition up to multiplication on the left by unitaries; these
3218: unitaries can always be uniquely chosen so as to satisfy
3219: \eqref{3.4a}.
3220: 
3221: Now define
3222: $$
3223: \rho_n^L = \kappa_n^L (\kappa_{n+1}^L)^{-1} \quad \text{and} \quad
3224: \rho_n^R =
3225: (\kappa_{n+1}^R)^{-1} \kappa_n^R.
3226: $$
3227: Notice that as inverses of positives matrices, $\rho_n^L >0$ and $
3228: \rho_n^R >0$.
3229: In particular, we have that
3230: $$
3231: \kappa_n^L = (\rho_{n-1}^L \dots \rho_0^L)^{-1} \quad \text{and}
3232: \quad \kappa_n^R =
3233: (\rho_0^R \dots \rho_{n-1}^R)^{-1}.
3234: $$
3235: 
3236: \begin{theorem}[Szeg\H{o} Recursion]\label{szethm}
3237: \begin{SL}
3238: \item[{\rm{(a)}}] For suitable matrices $\alpha_n^{L,R}$, one has
3239: \begin{align}
3240: z \varphi_n^L - \rho_n^L \varphi_{n+1}^L & = (\alpha_n^L)^\dagger
3241: \varphi_n^{R,*},  \lb{3.5a}\\
3242: z \varphi_n^R - \varphi_{n+1}^R \rho_n^R & = \varphi_n^{L,*}
3243: (\alpha_n^R)^\dagger. \lb{3.5b}
3244: \end{align}
3245: 
3246: \item[{\rm{(b)}}] The matrices $\alpha_n^L$ and $\alpha_n^R$ are
3247: equal and will henceforth be
3248: denoted by $\alpha_n$.\\
3249: 
3250: \item[{\rm{(c)}}] $\rho_n^L = (\bdone - \alpha_n^\dagger \alpha_n)^
3251: {1/2}$ and $\rho_n^R = (\bdone -
3252: \alpha_n \alpha_n^\dagger)^{1/2}$.
3253: \end{SL}
3254: \end{theorem}
3255: 
3256: \begin{proof}
3257: (a) The matrix polynomial $z \varphi_n^L$ has leading term $z^{n+1}
3258: \kappa_n^L$. On the
3259: other hand, the matrix polynomial $\rho_n^L \varphi_{n+1}^L$ has
3260: leading term $z^{n+1}
3261: \rho_n^L \kappa_{n+1}^L$. By definition of $\rho_n^L$, these terms
3262: are equal.
3263: Consequently, $z \varphi_n^L - \rho_n^L \varphi_{n+1}^L$ is a matrix
3264: polynomial of degree
3265: at most $n$. Notice that it is left-orthogonal with respect to $z
3266: \bdone, \ldots , z^n \bdone$ since
3267: $$
3268: \ang{z \varphi_n^L - \rho_n^L \varphi_{n+1}^L , z^j \bdone}_L =
3269: \ang{\varphi_n^L , z^{j-1} \bdone}_L - \ang{\rho_n^L \varphi_{n+1}
3270: ^L , z^j \bdone}_L
3271: = \bdzero - \bdzero = \bdzero.
3272: $$
3273: Now apply Lemma~\ref{l.c1}. The other claim is proved in the same way.\\
3274: 
3275: \smallskip
3276: (b) By part (a) and identities established earlier,
3277: \begin{align*}
3278: (\alpha_n^L)^\dagger & = \bdzero + (\alpha_n^L)^\dagger \bdone \\
3279: & = \ang{\varphi_n^{R,*} , \rho_n^L \varphi_{n+1}^L}_L +
3280: (\alpha_n^L)^\dagger \ang{\varphi_n^R ,  \varphi_n^R} _R \\
3281: & = \ang{\varphi_n^{R,*} , \rho_n^L \varphi_{n+1}^L}_L +
3282: (\alpha_n^L)^\dagger \ang{\varphi_n^{R,*} ,\varphi_n^{R,*}}_L \quad
3283: \text{by \eqref{3.1b}} \\
3284: & = \ang{\varphi_n^{R,*} , \rho_n^L \varphi_{n+1}^L +
3285: (\alpha_n^L)^\dagger \varphi_n^{R,*}}_L \\
3286: & = \ang{\varphi_n^{R,*} , z \varphi_n^L}_L \\
3287: & = \ang{z \varphi_n^R , \varphi_n^{L,*}}_R^\dagger  \quad \text
3288: {(using the $(n+1)$-degree *)}\\
3289: & = \ang{\varphi_{n+1}^R \rho_n^R + \varphi_n^{L,*} (\alpha_n^R)^
3290: \dagger ,
3291: \varphi_n^{L,*}}_R^\dagger \\
3292: & = \ang{\varphi_{n+1}^R \rho_n^R , \varphi_n^{L,*}}_R^\dagger
3293: + \ang{\varphi_n^{L,*} (\alpha_n^R)^\dagger , \varphi_n^{L,*}}_R^
3294: \dagger \\
3295: & = \bdzero + \ang{\varphi_n^{L,*} , \varphi_n^{L,*}(\alpha_n^R)^
3296: \dagger}_R \\
3297: & = \ang{\varphi_n^{L,*} , \varphi_n^{L,*}}_R \, (\alpha_n^R)^\dagger \\
3298: & = \ang{\varphi_n^L , \varphi_n^L}_L \,
3299: (\alpha_n^R)^\dagger \\
3300: & = (\alpha_n^R)^\dagger.
3301: \end{align*}
3302: (c) Using parts (a) and (b), we see that
3303: \begin{align*}
3304: \bdone & = \ang{z \varphi_n^L , z \varphi_n^L}_L \\
3305: & = \ang{\rho_n^L \varphi_{n+1}^L + \alpha_n^\dagger \varphi_n^{R,*} ,
3306: \rho_n^L \varphi_{n+1}^L + \alpha_n^\dagger \varphi_n^{R,*}}_L \\
3307: & = \rho_n^L \ang{\varphi_{n+1}^L , \varphi_{n+1}^L}_L
3308: (\rho_n^L)^\dagger + \alpha_n^\dagger \ang{\varphi_n^{R,*} ,
3309: \varphi_n^{R,*}}_L \, \alpha_n \\
3310: & = (\rho_n^L)2 + \alpha_n^\dagger \ang{\varphi_n^R , \varphi_n^R}_R
3311: \, \alpha_n \\
3312: & = (\rho_n^L)2 + \alpha_n^\dagger \alpha_n.
3313: \end{align*}
3314: A similar calculation yields the other claim.
3315: \end{proof}
3316: 
3317: The matrices $\alpha_n$ will henceforth be called the Verblunsky
3318: coefficients associated
3319: with the measure $d\mu$. Since $\rho_n^L$ is invertible, we have
3320: \begin{equation} \lb{3.3a}
3321: \norm{\alpha_n}<1.
3322: \end{equation}
3323: We will eventually see (Theorem~\ref{T3.10B}) that any set of $
3324: \alpha_n$'s obeying \eqref{3.3a}
3325: occurs as the set of Verblunsky coefficients for a unique non-trivial
3326: measure.
3327: 
3328: 
3329: Note that the Szeg\H{o} recursion for the monic orthogonal
3330: polynomials is
3331: \begin{equation} \label{szmonic}
3332: \begin{aligned}
3333: z \Phi_n^L -  \Phi_{n+1}^L & = (\kappa_n^L)^{-1} \alpha_n^\dagger
3334: (\kappa_n^R)^\dagger\Phi_n^{R,*}, \\
3335: z \Phi_n^R - \Phi_{n+1}^R  & = \Phi_n^{L,*} (\kappa_n^L)^\dagger
3336: \alpha_n^\dagger (\kappa_n^R)^{-1},
3337: \end{aligned}
3338: \end{equation}
3339: so the coefficients in the $L$ and $R$ equations are not equal and
3340: depend on all the
3341: $\alpha_j$, $j=1,\dots, n$.
3342: 
3343: Let us write the Szeg\H{o} recursion in matrix form, starting with
3344: left-orthogonal polynomials. By Theorem \ref{szethm},
3345: \begin{align*}
3346: \varphi_{n+1}^L & = (\rho_n^L)^{-1} [z \varphi_n^L - \alpha_n^\dagger
3347: \varphi_n^{R,*} ], \\
3348: \varphi_{n+1}^R & = [ z \varphi_n^R - \varphi_n^{L,*} \alpha_n^\dagger ]
3349: ( \rho_n^R )^{-1},
3350: \end{align*}
3351: which implies that
3352: \begin{align}
3353: \varphi_{n+1}^L & = z (\rho_n^L)^{-1} \varphi_n^L - (\rho_n^L)^{-1}
3354: \alpha_n^\dagger \varphi_n^{R,*}, \lb{3.7a} \\
3355: \varphi_{n+1}^{R,*} & = (\rho_n^R)^{-1} \varphi_n^{R,*} - z (\rho_n^R)
3356: ^{-1}
3357: \alpha_n \varphi_n^L. \lb{3.7b}
3358: \end{align}
3359: In other words,
3360: \begin{equation}
3361: \begin{pmatrix} \varphi_{n+1}^L \\ \varphi_{n+1}^{R,*} \end{pmatrix}
3362: = A^L(\alpha_n,z) \begin{pmatrix} \varphi_n^L \\ \varphi_n^{R,*} \end
3363: {pmatrix}
3364: \label{*1}
3365: \end{equation}
3366: where
3367: $$
3368: A^L(\alpha,z) = \begin{pmatrix} z (\rho^L)^{-1} & - (\rho^L)^{-1}
3369: \alpha^\dagger \\
3370: - z (\rho^R)^{-1} \alpha & (\rho^R)^{-1}
3371: \end{pmatrix}
3372: $$
3373: and $\rho^L = (\bdone - \alpha^\dagger \alpha)^{1/2}$, $\rho^R =
3374: (\bdone - \alpha
3375: \alpha^\dagger)^{1/2}$. Note that, for $z \not= 0$, the inverse of
3376: $A^L(\alpha,z)$ is
3377: given by
3378: $$
3379: A^L(\alpha,z)^{-1} =
3380: \begin{pmatrix}
3381: z^{-1} (\rho^L)^{-1} & z^{-1} (\rho^L)^{-1} \alpha^\dagger \\
3382: (\rho^R)^{-1} \alpha & (\rho^R)^{-1}
3383: \end{pmatrix}
3384: $$
3385: which gives rise to the inverse Szeg\H{o} recursion (first
3386: emphasized in the scalar and matrix cases by Delsarte el al.\
3387: \cite{DGK})
3388: \begin{align*}
3389: \varphi_n^L & =  z^{-1} (\rho_n^L)^{-1} \varphi_{n+1}^L + z^{-1}
3390: (\rho_n^L)^{-1} \alpha_n^\dagger \varphi_{n+1}^{R,*},
3391: \\
3392: \varphi_n^{R,*} & =  (\rho_n^R)^{-1} \alpha_n\varphi_{n+1}^L +
3393: (\rho_n^R)^{-1} \varphi_{n+1}^{R,*}.
3394: \end{align*}
3395: 
3396: For right-orthogonal polynomials, we find the following matrix
3397: formulas. By Theorem~\ref{szethm},
3398: \begin{align}
3399: \varphi_{n+1}^R & = z \varphi_n^R (\rho_n^R)^{-1} - \varphi_n^{L,*}
3400: \alpha_n^\dagger (\rho_n^R)^{-1},  \label{szright1}
3401: \\
3402: \varphi_{n+1}^{L,*} & = \varphi_n^{L,*} (\rho_n^L)^{-1} - z \varphi_n^R
3403: \alpha_n (\rho_n^L)^{-1}.  \label{szright2}
3404: \end{align}
3405: In other words,
3406: $$
3407: \begin{pmatrix} \varphi_{n+1}^R & \varphi_{n+1}^{L,*} \end{pmatrix}  =
3408: \begin{pmatrix} \varphi_n^R & \varphi_n^{L,*} \end{pmatrix} A^R
3409: (\alpha_n,z)
3410: $$
3411: where
3412: $$
3413: A^R(\alpha,z) = \begin{pmatrix}
3414: z (\rho^R)^{-1} & - z \alpha (\rho^L)^{-1} \\
3415: - \alpha^\dagger (\rho^R)^{-1} & (\rho^L)^{-1}
3416: \end{pmatrix}.
3417: $$
3418: For $z \not= 0$, the inverse of $A^R(\alpha,z)$ is given by
3419: $$
3420: A^R(\alpha,z)^{-1} = \begin{pmatrix}
3421: z^{-1} (\rho^R)^{-1} & (\rho^R)^{-1} \alpha \\
3422: z^{-1} (\rho^L)^{-1} \alpha^\dagger & (\rho^L)^{-1}
3423: \end{pmatrix}
3424: $$
3425: and hence
3426: \begin{align}
3427: \varphi_n^R & =  z^{-1} \varphi_{n+1}^R (\rho_n^R)^{-1} + z^{-1}
3428: \varphi_{n+1}^{L,*} (\rho_n^L)^{-1} \alpha_n^\dagger,  \label{szright3}
3429: \\
3430: \varphi_n^{L,*} & =  \varphi_{n+1}^R (\rho_n^R)^{-1} \alpha_n +
3431: \varphi_{n+1}^{L,*} (\rho_n^L)^{-1}. \label{szright4}
3432: \end{align}
3433: 
3434: 
3435: 
3436: \subsection{Second Kind Polynomials} \lb{s3.3}
3437: 
3438: In the scalar case, second kind polynomials go back to Geronimus \cite
3439: {Ger44,GBk1,GBk}.
3440: For $n\geq1$, let us introduce the second kind polynomials $\psi_n^
3441: {L,R}$ by
3442: \begin{align}
3443: \psi_n^L (z)&= \int \frac{e^{i\theta}+z}{e^{i\theta}-z}\,
3444: (\varphi_n^L ( e^{i\theta} )
3445: -\varphi_n^L (z) ) \, d\mu (\theta), \label{a1}
3446: \\
3447: \psi_n^R (z)&= \int \frac{e^{i\theta}+z}{e^{i\theta}-z} \, d\mu
3448: (\theta) \, (\varphi_n^R
3449: ( e^{i\theta} ) -\varphi_n^R (z) ) . \label{a2}
3450: \end{align}
3451: For $n=0$, let us set $\psi_0^L (z)=\psi_0^R (z)=\bdone$. For
3452: future reference, let us display the first polynomials of each
3453: series:
3454: \begin{align}
3455: \varphi_1^L(z)&=(\rho_0^L)^{-1}(z-\alpha_0^\dagger),
3456: \qquad
3457: \varphi_1^R(z)=(z-\alpha_0^\dagger)(\rho_0^R)^{-1},
3458: \label{firstpoly1}
3459: \\
3460: \psi_1^L(z)&=(\rho_0^L)^{-1}(z+\alpha_0^\dagger),
3461: \qquad
3462: \psi_1^R(z)=(z+\alpha_0^\dagger)(\rho_0^R)^{-1}.
3463: \label{firstpoly2}
3464: \end{align}
3465: We will also need formulas for
3466: $\psi_n^{L,*}$ and $\psi_n^{R,*}$, $n\geq1$. These formulas follow
3467: directly from the
3468: above definition and from
3469: $$
3470: \overline{\left(\frac{e^{i\theta}+1/\bar z}{e^{i\theta}-1/\bar z}
3471: \right)} =
3472: -\frac{e^{i\theta}+z}{e^{i\theta}-z}\, .
3473: $$
3474: Indeed, we have
3475: \begin{align}
3476: \psi_n^{L,*} (z)&= z^n \int \frac{e^{i\theta}+z}{e^{i\theta}-z} \, d
3477: \mu (\theta) \,
3478: (\varphi_n^L (1/\bar z )^\dagger -\varphi_n^L (e^{i\theta})^\dagger),
3479: \label{a2a}
3480: \\
3481: \psi_n^{R,*} (z)&= z^n \int \frac{e^{i\theta}+z}{e^{i\theta}-z}\,
3482: (\varphi_n^R
3483: (1/\bar z )^\dagger -\varphi_n^R (e^{i\theta})^\dagger) \, d\mu
3484: (\theta).
3485: \label{a3}
3486: \end{align}
3487: 
3488: \begin{proposition}
3489: The second kind polynomials obey the recurrence relations
3490: \begin{align}
3491: \psi_{n+1}^L(z)&=(\rho_n^L)^{-1}(z\psi_n^L(z)+\alpha_n^\dagger \psi_
3492: {n}^{R,*}(z)),
3493: \label{a4}
3494: \\
3495: \psi_{n+1}^{R,*}(z)&=(\rho_n^R)^{-1}(z\alpha_n \psi_{n}^L(z)+\psi_{n}^
3496: {R,*}(z))
3497: \label{a5}
3498: \end{align}
3499: and
3500: \begin{align}
3501: \psi_{n+1}^R(z) & = (z \psi_n^R(z) + \psi_n^{L,*}(z) \alpha_n^
3502: \dagger) (\rho_n^R)^{-1}, \label{a4r}
3503: \\
3504: \psi_{n+1}^{L,*}(z) & = (\psi_n^{L,*}(z) + z \psi_n^R(z) \alpha_n)
3505: (\rho_n^L)^{-1} \label{a5r}
3506: \end{align}
3507: for $n\geq0$.
3508: \end{proposition}
3509: \begin{proof}
3510: 1. Let us check \eqref{a4} for $n\geq1$. Denote the right-hand side
3511: of \eqref{a4} by
3512: $\tilde\psi_{n+1}^L(z)$. Using the recurrence relations for $
3513: \varphi_n^L$, $\varphi_n^{R,*}$ and
3514: the definition \eqref{a1} of $\psi_n^L$, we find
3515: $$
3516: \psi_{n+1}^L(z)- \tilde \psi_{n+1}^L(z) = \int \frac{e^{i\theta}+z}{e^
3517: {i\theta}-z} \,
3518: A_n(\theta,z) \, d\mu(\theta)
3519: $$
3520: where
3521: \begin{align*}
3522: A_n(\theta,z) & = \varphi_{n+1}^L(e^{i\theta}) -\varphi_{n+1}^L(z)
3523: \\
3524: & \qquad - (\rho_n^L)^{-1}[z \varphi_{n}^L(e^{i\theta}) -z\varphi_{n}
3525: ^L(z) +
3526: \alpha_n^\dagger z^n \varphi_n^R(1/\bar z)^\dagger
3527: - \alpha_n^\dagger z^n \varphi_n^R(e^{i\theta})^\dagger]
3528: \\
3529: & = (\rho_n^L)^{-1} [ e^{i\theta}\varphi_{n}^L(e^{i\theta}) -
3530: \alpha_n^\dagger\varphi_{n}^{R,*}(e^{i\theta}) -
3531: z\varphi_{n}^L(z)+\alpha_n^\dagger\varphi_{n}^{R,*}(z)
3532: \\
3533: & \qquad - z \varphi_{n}^L(e^{i\theta}) +z\varphi_{n}^L(z) - \alpha_n^
3534: \dagger
3535: \varphi_n^{R,*}(z) + \alpha_n^\dagger z^n
3536: \varphi_n^R(e^{i\theta})^\dagger ]
3537: \\
3538: & = (\rho_n^L)^{-1}[(e^{i\theta}-z)\varphi_{n}^L(e^{i\theta}) +
3539: \alpha_n^\dagger(z^n
3540: e^{-in\theta}-1)\varphi_n^{R,*}(e^{i\theta})].
3541: \end{align*}
3542: Using the orthogonality relations
3543: \begin{equation}
3544: \int \varphi_n^L(e^{i\theta})\, d\mu(\theta) e^{-im\theta} = \int
3545: \varphi_n^{R,*}(e^{i\theta})\, d\mu(\theta) e^{-i(m+1)\theta} =
3546: \bdzero, \label{a6}
3547: \end{equation}
3548: $m=0,1,\dots,n-1$, and the formula
3549: $$
3550: \frac{e^{in\theta}-z^n}{e^{i\theta}-z} =
3551: e^{i(n-1)\theta}+e^{i(n-2)\theta}z+\dots+z^{n-1}
3552: $$
3553: we obtain
3554: \begin{align*}
3555: \rho_n^L \int \frac{e^{i\theta}+z}{e^{i\theta}-z}\, A_n(\theta,z)\, d
3556: \mu(\theta) & = \int
3557: (e^{i\theta}\varphi_n^L(e^{i\theta}) - \alpha_n^\dagger
3558: \varphi_n^{R,*}(e^{i\theta}))\, d\mu(\theta) \\
3559: & = \rho_n^L\int \varphi_{n+1}^L (e^{i\theta}) \, d\mu(\theta) =
3560: \bdzero.
3561: \end{align*}
3562: 
3563: \smallskip
3564: 2. Let us check \eqref{a5} for $n\geq1$. Denote the right-hand side
3565: of \eqref{a5} by
3566: $\tilde\psi_{n+1}^{R,*}(z)$. Similarly to the argument above, we find
3567: $$
3568: \psi_{n+1}^{R,*}(z)- \tilde \psi_{n+1}^{R,*}(z) = \int
3569: \frac{e^{i\theta}+z}{e^{i\theta}-z} \, B_n(\theta,z) \, d\mu(\theta),
3570: $$
3571: where
3572: \begin{align*}
3573: B_n(\theta,z) & = z^{n+1}\varphi_{n+1}^R(1/\bar z)^\dagger -
3574: z^{n+1}\varphi_{n+1}^R(e^{i\theta})^\dagger \\
3575: & \qquad - (\rho_n^R)^{-1} [ z\alpha_n\varphi_n^L(e^{i\theta}) - z
3576: \alpha_n\varphi_n^L(z)
3577: +
3578: z^n\varphi_n^R(1/\bar z)^\dagger - z^n\varphi_n^R(e^{i\theta})^
3579: \dagger ] \\
3580: & = (\rho_n^R)^{-1} [ \varphi_n^{R,*}(z) - \alpha_n z \varphi_n^L(z) -
3581: z^{n+1}e^{-i(n+1)\theta}
3582: \varphi_n^{R,*}(e^{i\theta}) + z^{n+1}e^{-in\theta}\alpha_n
3583: \varphi_n^L(e^{i\theta}) \\
3584: & \qquad - z \alpha_n \varphi_n^L(e^{i\theta}) + z \alpha_n
3585: \varphi_n^L(z) -
3586: \varphi_n^{R,*}(z) +
3587: z^ne^{-in\theta} \varphi_n^{R,*}(e^{i\theta}) ] \\
3588: & = (\rho_n^R)^{-1} [ z \alpha_n (z^n e^{-in\theta}-1) \varphi_n^L(e^
3589: {i\theta}) + z^n
3590: e^{-in\theta}(1-ze^{-i\theta})\varphi_n^{R,*}(e^{i\theta}) ].
3591: \end{align*}
3592: Using the orthogonality relations \eqref{a6}, we get
3593: \begin{align*}
3594: \rho_n^R \int \frac{e^{i\theta}+z}{e^{i\theta}-z} \, & B_n(\theta,z)
3595: \, d\mu(\theta) = \\
3596: & = z^{n+1}\int ( e^{-i(n+1)\theta}\varphi_n^{R,*}(e^{i\theta}) -
3597: \alpha_n
3598: e^{-in\theta}\varphi_n^{L}(e^{i\theta})) \, d\mu(\theta) \\
3599: & = z^{n+1}\rho_n^R \int e^{-i(n+1)\theta} \varphi_{n+1}^{R,*}(e^{i
3600: \theta}) \, d\mu(\theta) = \bdzero.
3601: \end{align*}
3602: 
3603: \smallskip
3604: 3. Relations \eqref{a4} and \eqref{a5} can be checked for $n=0$ by
3605: a direct substitution of \eqref{firstpoly1} and
3606: \eqref{firstpoly2}.
3607: 
3608: \smallskip
3609: 4. We obtain \eqref{a4r} and \eqref{a5r} from \eqref{a4} and \eqref
3610: {a5} by applying the
3611: $*$-operation.
3612: \end{proof}
3613: 
3614: Writing the above recursion in matrix form, we get
3615: $$
3616: \begin{pmatrix} \psi_{n+1}^L \\ \psi_{n+1}^{R,*} \end{pmatrix}
3617: = A^L(-\alpha_n,z) \begin{pmatrix} \psi_n^L \\ \psi_n^{R,*} \end
3618: {pmatrix}, \qquad
3619: \begin{pmatrix} \psi_0^L \\ \psi_0^{R,*} \end{pmatrix} =
3620: \begin{pmatrix} \bdone \\ \bdone \end{pmatrix}
3621: $$
3622: for left-orthogonal polynomials and
3623: $$
3624: \begin{pmatrix} \psi_{n+1}^R & \psi_{n+1}^{L,*} \end{pmatrix} =
3625: \begin{pmatrix} \psi_n^R & \psi_n^{L,*} \end{pmatrix} A^R(-\alpha_n,z),
3626: \qquad
3627: \begin{pmatrix} \psi_0^R & \psi_0^{L,*} \end{pmatrix} =
3628: \begin{pmatrix} \bdone & \bdone \end{pmatrix}.
3629: $$
3630: for right-orthogonal polynomials.
3631: 
3632: Equivalently,
3633: \begin{equation}
3634: \begin{pmatrix} \psi_{n+1}^L \\ - \psi_{n+1}^{R,*} \end{pmatrix}
3635: = A^L(\alpha_n,z) \begin{pmatrix} \psi_n^L \\ - \psi_n^{R,*} \end
3636: {pmatrix}, \qquad
3637: \begin{pmatrix} \psi_0^L \\ - \psi_0^{R,*} \end{pmatrix} =
3638: \begin{pmatrix} \bdone \\ - \bdone \end{pmatrix}
3639: \label{*2}
3640: \end{equation}
3641: and
3642: $$
3643: \begin{pmatrix} \psi_{n+1}^R & - \psi_{n+1}^{L,*} \end{pmatrix} =
3644: \begin{pmatrix} \psi_n^R & - \psi_n^{L,*} \end{pmatrix} A^R
3645: (\alpha_n,z), \qquad
3646: \begin{pmatrix} \psi_0^R & - \psi_0^{L,*} \end{pmatrix} =
3647: \begin{pmatrix} \bdone & - \bdone \end{pmatrix}.
3648: $$
3649: 
3650: In particular, we see that the second kind polynomials $\psi_n^{L,R}$
3651: correspond to
3652: Verblunsky coefficients $\{ - \alpha_n \}$. We have the following
3653: Wronskian-type relations:
3654: 
3655: \begin{proposition}
3656: For $n \ge 0$ and $z \in \C$, we have
3657: \begin{align}
3658: 2z^n \bdone & = \varphi_n^L(z) \psi_n^{L,*}(z) + \psi_n^L(z)
3659: \varphi_n^{L,*}(z), \label{wronsk1} \\
3660: 2z^n \bdone & = \psi_n^{R,*}(z) \varphi_n^R(z) + \varphi_n^{R,*}(z)
3661: \psi_n^R(z). \label{wronsk2} \\
3662: \bdzero & = \varphi_n^L(z) \psi_n^R(z) - \psi_n^L(z) \varphi_n^R(z),
3663: \label{wronsk3} \\
3664: \bdzero & = \psi_n^{R,*}(z) \varphi_n^{L,*}(z) - \varphi_n^{R,*}(z)
3665: \psi_n^{L,*}(z).
3666: \label{wronsk4}
3667: \end{align}
3668: \end{proposition}
3669: 
3670: \begin{proof}
3671: We prove this by induction. The four identities clearly hold for
3672: $n = 0$. Suppose \eqref{wronsk1}--\eqref{wronsk4} hold for some $n
3673: \ge 0$. Then,
3674: \begin{align*}
3675: \varphi_{n+1}^L \psi_{n+1}^{L,*} & + \psi_{n+1}^L \varphi_{n+1}^{L,*} =
3676: \\
3677: & =
3678: \left(\rho_n^L\right)^{-1} [(z \varphi_n^L - \alpha_n^\dagger
3679: \varphi_n^{R,*})
3680: (\psi_n^{L,*} + z \psi_n^R \alpha_n)
3681: \\
3682: & \qquad + (z\psi_n^L+\alpha_n^\dagger \psi_{n}^{R,*})
3683: (\varphi_n^{L,*} - z \varphi_n^R \alpha_n) ] \left(\rho_n^L\right)^{-1}
3684: \\
3685: & = \left(\rho_n^L\right)^{-1} [ z(\varphi_n^L \psi_n^{L,*} +
3686: \psi_n^L \varphi_n^{L,*}) -
3687: z \alpha_n^\dagger ( \psi_n^{R,*} \varphi_n^R + \varphi_n^{R,*}
3688: \psi_n^R )\alpha_n
3689: \\
3690: & \qquad + \alpha_n^\dagger ( \psi_n^{R,*} \varphi_n^{L,*} -
3691: \varphi_n^{R,*} \psi_n^{L,*}
3692: ) + z2 ( \varphi_n^L \psi_n^R - \psi_n^L \varphi_n^R ) \alpha_n ]
3693: \left(\rho_n^L\right)^{-1}
3694: \\
3695: & = \left(\rho_n^L\right)^{-1} [ 2z^{n+1} (\bdone - \alpha_n^\dagger
3696: \alpha_n) ]
3697: \left(\rho_n^L\right)^{-1} = 2z^{n+1} \bdone,
3698: \end{align*}
3699: where we used \eqref{wronsk1}--\eqref{wronsk4} for $n$ in the third
3700: step. Thus,
3701: \eqref{wronsk1} holds for $n+1$.
3702: 
3703: For \eqref{wronsk2}, we note that
3704: \begin{align*}
3705: \psi_{n+1}^{R,*} \varphi_{n+1}^R & + \varphi_{n+1}^{R,*} \psi_{n+1}^R =
3706: \\
3707: & = (\rho_n^R)^{-1} [
3708: (z\alpha_n \psi_{n}^L + \psi_{n}^{R,*}) (z \varphi_n^R - \varphi_n^{L,*}
3709: \alpha_n^\dagger)
3710: \\
3711: & \qquad + (\varphi_n^{R,*} - z \alpha_n \varphi_n^L)(z \psi_n^R+
3712: \psi_n^{L,*}
3713: \alpha_n^\dagger)]
3714: \left( \rho_n^R \right)^{-1}
3715: \\
3716: & = (\rho_n^R)^{-1} [ z (\psi_n^{R,*} \varphi_n^R + \varphi_n^{R,*}
3717: \psi_n^R) - z
3718: \alpha_n (\psi_n^L \varphi_n^{L,*} + \varphi_n^L \psi_n^{L,*})
3719: \alpha_n^\dagger
3720: \\
3721: & \qquad + z2 \alpha_n (\psi_n^L \varphi_n^R - \varphi_n^L \psi_n^R)
3722: - (\psi_n^{R,*}
3723: \varphi_n^{L,*} - \varphi_n^{R,*} \psi_n^{L,*}) \alpha_n^\dagger ]
3724: \left( \rho_n^R \right)^{-1}
3725: \\
3726: & = (\rho_n^R)^{-1} 2 z^{n+1}( \bdone - \alpha_n \alpha_n^\dagger )
3727: \left( \rho_n^R
3728: \right)^{-1} = 2z^{n+1} \bdone ,
3729: \end{align*}
3730: again using \eqref{wronsk1}--\eqref{wronsk4} for $n$ in the third step.
3731: 
3732: Next,
3733: \begin{align*}
3734: \varphi_{n+1}^L \psi_{n+1}^R & - \psi_{n+1}^L \varphi_{n+1}^R = \\
3735: & = (\rho_n^L)^{-1} [ (z
3736: \varphi_n^L - \alpha_n^\dagger \varphi_n^{R,*}) (z\psi_n^R + \psi_n^
3737: {L,*} \alpha_n^\dagger) \\
3738: & \qquad - (z\psi_n^L + \alpha_n^\dagger \psi_n^{R,*}) (z\varphi_n^R
3739: - \varphi_n^{L,*}
3740: \alpha_n^\dagger) ] (\rho_n^R)^{-1} \\
3741: & = (\rho_n^L)^{-1} [ z2(\varphi_n^L \psi_n^R - \psi_n^L
3742: \varphi_n^R) - \alpha_n^\dagger
3743: (\varphi_n^{R,*} \psi_n^{L,*} - \psi_n^{R,*} \varphi_n^{L,*})
3744: \alpha_n^\dagger \\
3745: & \qquad - z \alpha_n^\dagger (\varphi_n^{R,*} \psi_n^R + \psi_n^
3746: {R,*} \varphi_n^R) + z
3747: (\varphi_n^L \psi_n^{L,*} + \psi_n^L \varphi_n^{L,*}) \alpha_n^
3748: \dagger ] (\rho_n^R)^{-1}
3749: \\
3750: &= \bdzero
3751: \end{align*}
3752: which implies first \eqref{wronsk3} for $n+1$ and then, by applying
3753: the $*$-operation of
3754: order $2n+2$, also \eqref{wronsk4} for $n+1$. This concludes the
3755: proof of the
3756: proposition.
3757: \end{proof}
3758: 
3759: \subsection{Christoffel--Darboux Formulas} \lb{s3.3A}
3760: 
3761: \begin{proposition} \lb{P3.6}
3762: {\rm (a) (CD)-left orthogonal}
3763: \begin{align*}
3764: (1 - \bar \xi z) \sum_{k = 0}^n \varphi_k^L (\xi)^\dagger \varphi_k^L
3765: (z)
3766: & = \varphi_n^{R,*} (\xi)^\dagger \varphi_n^{R,*} (z) - \bar \xi z
3767: \varphi_n^L (\xi)^\dagger \varphi_n^L (z)
3768: \\
3769: & = \varphi_{n+1}^{R,*} (\xi)^\dagger \varphi_{n+1}^{R,*} (z) -
3770: \varphi_{n+1}^L (\xi)^\dagger \varphi_{n+1}^L (z).
3771: \end{align*}
3772: {\rm (b) (CD)-right orthogonal}
3773: \begin{align*}
3774: (1 - \bar \xi z) \sum_{k = 0}^n \varphi_k^R (z) \varphi_k^R (\xi)^
3775: \dagger
3776: & = \varphi_n^{L,*} (z) \varphi_n^{L,*} (\xi)^\dagger - \bar \xi z
3777: \varphi_n^R (z) \varphi_n^R (\xi)^\dagger \\
3778: & = \varphi_{n+1}^{L,*} (z) \varphi_{n+1}^{L,*} (\xi)^\dagger -
3779: \varphi_{n+1}^R (z) \varphi_{n+1}^R (\xi)^\dagger.
3780: \end{align*}
3781: {\rm (c) (Mixed CD)-left orthogonal}
3782: \begin{align*}
3783: (1 - \bar \xi z) \sum_{k = 0}^n \psi_k^L (\xi)^\dagger
3784: \varphi_k^L (z) & = 2\cdot \bdone - \psi_n^{R,*} (\xi)^\dagger
3785: \varphi_n^{R,*} (z) - \bar \xi z
3786: \psi_n^L (\xi)^\dagger \varphi_n^L (z) \\
3787: & = 2\cdot \bdone - \psi_{n+1}^{R,*} (\xi)^\dagger \varphi_{n+1}^
3788: {R,*} (z) -
3789: \psi_{n+1}^L (\xi)^\dagger \varphi_{n+1}^L (z).
3790: \end{align*}
3791: {\rm (d) (Mixed CD)-right orthogonal}
3792: \begin{align*}
3793: (1 - \bar \xi z) \sum_{k = 0}^n \varphi_k^R (z) \psi_k^R (\xi)^
3794: \dagger & =
3795: 2\cdot\bdone - \varphi_n^{L,*} (z) \psi_n^{L,*} (\xi)^\dagger - \bar
3796: \xi z
3797: \varphi_n^R (z) \psi_n^R (\xi)^\dagger \\
3798: & = 2\cdot\bdone - \varphi_{n+1}^{L,*} (z) \psi_{n+1}^{L,*} (\xi)^
3799: \dagger -
3800: \varphi_{n+1}^R (z) \psi_{n+1}^R (\xi)^\dagger.
3801: \end{align*}
3802: \end{proposition}
3803: 
3804: \begin{remark} Since the $\psi$'s are themselves MOPUCs, the analogue of
3805: (a) and (b), with all $\varphi$'s replaced by $\psi$'s, holds.
3806: \end{remark}
3807: 
3808: \begin{proof}
3809: (a) Write
3810: $$
3811: F^L_n(z) = \begin{pmatrix} \varphi_n^L(z) \\
3812: \varphi_n^{R,*}(z)\end {pmatrix}, \quad J = \left(
3813: \begin{array}{rr} \bdone & \bdzero \\ \bdzero & -\bdone
3814: \end{array} \right), \quad
3815: \tilde J  = \left( \begin{array}{cr} \bar \xi z \bdone & \bdzero \\
3816: \bdzero & -\bdone
3817: \end{array} \right).
3818: $$
3819: Then,
3820: $$
3821: F^L_{n+1}(z) = A^L (\alpha_n,z) F^L_n(z)
3822: $$
3823: and
3824: \begin{align*}
3825: A^L (\alpha,\xi)^\dagger & J A^L(\alpha,z) = \\
3826: & = \begin{pmatrix} \bar \xi (\rho^L)^{-1} & - \bar \xi \alpha^
3827: \dagger (\rho^R)^{-1} \\
3828: -\alpha (\rho^L)^{-1} & (\rho^R)^{-1} \end{pmatrix}
3829: \begin{pmatrix} z (\rho^L)^{-1} & - (\rho^L)^{-1} \alpha^\dagger \\
3830: z (\rho^R)^{-1} \alpha & - (\rho^R)^{-1} \end{pmatrix} \\
3831: & = \begin{pmatrix} \bar \xi z (\rho^L)^{-2} - \bar \xi z
3832: \alpha^\dagger (\rho^R)^{-2} \alpha & - \bar \xi (\rho^L)^{-2}
3833: \alpha^\dagger + \bar \xi \alpha^\dagger (\rho^R)^{-2} \\
3834: - z \alpha (\rho^L)^{-2} + z (\rho^R)^{-2} \alpha & \alpha
3835: (\rho^L)^{-2} \alpha^\dagger - (\rho^R)^{-2} \end{pmatrix} \\
3836: & = \begin{pmatrix} \bar \xi z \bdone & \bdzero \\ \bdzero & - \bdone
3837: \end{pmatrix}
3838: = \tilde J .
3839: \end{align*}
3840: Thus,
3841: \begin{align*}
3842: F^L_{n+1}(\xi)^\dagger J F^L_{n+1}(z) &=  F^L_n(\xi)^\dagger A^L
3843: (\alpha_n,\xi)^\dagger J A^L
3844: (\alpha_n,z) F^L_n(z) \\
3845: &= F^L_n(\xi)^\dagger \tilde J F^L_n(z)
3846: \end{align*}
3847: and hence
3848: $$
3849: \varphi_{n+1}^L(\xi)^\dagger \varphi_{n+1}^L(z) - \varphi_{n+1}^{R,*}
3850: (\xi)^\dagger
3851: \varphi_{n+1}^{R,*}(z) = \bar \xi z \varphi_n^L(\xi)^\dagger
3852: \varphi_n^L(z) -
3853: \varphi_n^{R,*}(\xi)^\dagger \varphi_n^{R,*}(z)
3854: $$
3855: which shows that the last two expressions in (a) are equal. Denote
3856: their common value by
3857: $Q^L_n(z,\xi)$. Then,
3858: \begin{align*}
3859: Q^L_n  (z,\xi) - Q^L_{n-1}(z,\xi) & =
3860:     \varphi_n^{R,*} (\xi)^\dagger \varphi_n^{R,*} (z) - \bar \xi z
3861: \varphi_n^L (\xi)^\dagger \varphi_n^L (z) \\
3862: & \qquad - \varphi_n^{R,*} (\xi)^\dagger \varphi_n^{R,*} (z) +
3863: \varphi_n^L (\xi)^\dagger \varphi_n^L (z) \\
3864: & = (1 - \bar \xi z) \varphi_n^L (\xi)^\dagger \varphi_n^L (z).
3865: \end{align*}
3866: Summing over $n$ completes the proof since $Q^L_{-1}(z,\xi)=0$.
3867: 
3868: \smallskip
3869: (b) The proof is analogous to (a): Write $F^R_n(z) = \left(
3870: \varphi_n^R(z) \quad  \varphi_n^{L,*}(z) \right)$. Then,
3871: $F^R_{n+1}(z) = F^R_n(z) A^R(\alpha_n,z)$ and $A^R (\alpha,z) J
3872: A^R (\alpha,\xi)^ \dagger = \tilde J$. Thus,
3873: \begin{align*}
3874: F^R_{n+1}(z) J F^R_{n+1}(\xi)^\dagger & = F^R_n(z) A^R(\alpha_n,z)
3875: J A^R (\alpha_n,\xi)^\dagger F^R_n
3876: (\xi)^\dagger \\
3877: &= F^R_n(z) \tilde J F^R_n(\xi)^\dagger
3878: \end{align*}
3879: and hence
3880: $$
3881: \varphi_{n+1}^R(z) \varphi_{n+1}^R(\xi)^\dagger - \varphi_{n+1}^{L,*}(z)
3882: \varphi_{n+1}^{L,*}(\xi)^\dagger = \bar \xi z \varphi_n^R(z)
3883: \varphi_n^R(\xi)^\dagger -
3884: \varphi_n^{L,*}(z) \varphi_n^{L,*}(\xi)^\dagger
3885: $$
3886: which shows that the last two expressions in (b) are equal. Denote
3887: their common value by $Q^R_n(z,\xi)$. Then,
3888: $$
3889: Q^R_n (z,\xi) - Q^R_{n-1}(z,\xi) = (1 - \bar \xi z) \varphi_n^R
3890: (z) \varphi_n^R (\xi)^\dagger
3891: $$
3892: and the assertion follows as before.
3893: 
3894: \smallskip
3895: (c) Write
3896: $$
3897: \tilde F^L_n(z) =  \begin{pmatrix} \psi_n^L(z) \\ -
3898: \psi_n^{R,*}(z)
3899: \end{pmatrix}
3900: $$
3901: with the second kind polynomials $\psi_n^{L,R}$. As in (a), we see
3902: that
3903: \begin{align*}
3904: \tilde F^L_{n+1}(\xi)^\dagger J F^L_{n+1}(z) &=  \tilde
3905: F^L_n(\xi)^\dagger A^L(\alpha_n,\xi)^\dagger J A^L
3906: (\alpha_n,z) F^L_n(z) \\
3907: & = \tilde F^L_n(\xi)^\dagger \tilde J F^L_n(z)
3908: \end{align*}
3909: and hence
3910: $$
3911: \psi_{n+1}^L(\xi)^\dagger \varphi_{n+1}^L(z) + \psi_{n+1}^{R,*}(\xi)^
3912: \dagger
3913: \varphi_{n+1}^{R,*}(z) = \bar \xi z \psi_n^L(\xi)^\dagger \varphi_n^L
3914: (z) +
3915: \psi_n^{R,*}(\xi)^\dagger \varphi_n^{R,*}(z).
3916: $$
3917: Denote
3918: $$
3919: \tilde Q^L_n(z,\xi) = 2\cdot \bdone - \psi_{n+1}^{R,*}(\xi)^\dagger
3920: \varphi_{n+1}^{R,*}(z) -
3921: \psi_{n+1}^L(\xi)^\dagger \varphi_{n+1}^L(z).
3922: $$
3923: Then,
3924: \begin{align*}
3925: \tilde Q^L_n  (z,\xi) - \tilde Q^L_{n-1}(z,\xi)
3926: & = - \psi_n^{R,*} (\xi)^\dagger \varphi_n^{R,*} (z) - \bar \xi z
3927: \psi_n^L (\xi)^\dagger \varphi_n^L (z)\\
3928: & \qquad  + \psi_n^{R,*} (\xi)^\dagger \varphi_n^{R,*} (z) + \psi_n^L
3929: (\xi)^\dagger \varphi_n^L (z) \\
3930: & = (1 - \bar \xi z) \psi_n^L (\xi)^\dagger \varphi_n^L (z)
3931: \end{align*}
3932: and the assertion follows as before.
3933: 
3934: \smallskip
3935: (d) Write $\tilde F^R_n(z) = \left(  \psi_n^R(z) \quad -
3936: \psi_n^{L,*}(z) \right)$. As in (b), we see that
3937: \begin{align*}
3938: F^R_{n+1}(z) J \tilde F^R_{n+1}(\xi)^\dagger &= F^R_n(z)
3939: A^R(\alpha_n,z) J A^R(\alpha_n,\xi)^\dagger \tilde
3940: F^R_n(\xi)^\dagger \\
3941: &= F^R_n(z) \tilde J \tilde F^R_n(\xi)^\dagger
3942: \end{align*}
3943: and hence
3944: $$
3945: \varphi_{n+1}^R(z) \psi_{n+1}^R(\xi)^\dagger + \varphi_{n+1}^{L,*}(z)
3946: \psi_{n+1}^{L,*}(\xi)^\dagger = \bar \xi z \varphi_n^R(z) \psi_n^R
3947: (\xi)^\dagger +
3948: \varphi_n^{L,*}(z) \psi_n^{L,*}(\xi)^\dagger.
3949: $$
3950: With $\tilde Q^R_n(z,\xi) = 2\cdot\bdone - \varphi_{n+1}^{L,*}(z)
3951: \psi_{n +1}^{L,*}(\xi)^\dagger - \varphi_{n+1}^R(z)
3952: \psi_{n+1}^R(\xi)^\dagger$, we have
3953: $$
3954: \tilde Q^R_n (z,\xi) - \tilde Q^R_{n-1}(z,\xi) = (1 - \bar \xi z)
3955: \varphi_n^R (z)
3956: \psi_n^R (\xi)^\dagger
3957: $$
3958: and we conclude as in (c).
3959: \end{proof}
3960: 
3961: \subsection{Zeros of MOPUC} \lb{s3.3B}
3962: 
3963: Our main result in this section is:
3964: 
3965: \begin{theorem}\lb{T3.6A} All the zeros of $\det(\varphi_n^R(z))$ lie
3966: in $\bbD=
3967: \{z\, \colon \abs{z}<1\}$.
3968: \end{theorem}
3969: 
3970: We will also prove:
3971: 
3972: \begin{theorem}\lb{T3.6B} For each $n$,
3973: \begin{equation}\lb{3.24a}
3974: \det(\varphi_n^R(z))=\det(\varphi_n^L(z)).
3975: \end{equation}
3976: \end{theorem}
3977: 
3978: The scalar analogue of Theorem~\ref{T3.6A} has seven proofs in \cite
3979: {S}! The simplest is due to
3980: Landau \cite{Landau} and its MOPUC analogue is Theorem~2.13.7 of \cite
3981: {S}. There is also a proof
3982: in Delsarte et al.\ \cite{DGK} who attribute the theorem to Whittle
3983: \cite{Whit}.  We provide two
3984: more proofs here, not only for their intrinsic interest: our first
3985: proof we need because it
3986: depends only on the recursion relation (it is related to the proof of
3987: Delsarte et al.\ \cite{DGK}).
3988: The second proof is here since it relates zeros to eigenvalues of a
3989: cutoff CMV matrix.
3990: 
3991: \begin{theorem}\lb{T3.6C} We have
3992: \begin{SL}
3993: \item[{\rm{(i)}}] For $z\in\partial\bbD$, all of $\varphi_n^{R,*}(z)
3994: $, $\varphi_n^{L,*}(z)$, $\varphi_n^R(z)$,
3995: $\varphi_n^L(z)$ are invertible.
3996: 
3997: \item[{\rm{(ii)}}] For $z\in\partial\bbD$, $\varphi_n^L(z) (\varphi_n^
3998: {R,*}(z))^{-1}$ and $(\varphi_n^{*,L}(z))^{-1}
3999: \varphi_n^R(z)$ are unitary.
4000: 
4001: \item[{\rm{(iii)}}] For $z\in\bbD$, $\varphi_n^{R,*}(z)$ and
4002: $\varphi_n^{L,*}(z)$ are invertible.
4003: 
4004: \item[{\rm{(iv)}}] For $z\in\bbD$, $\varphi_n^L(z)(\varphi_n^{R,*}(z))
4005: ^{-1}$ and $(\varphi_n^{*,L}(z))^{-1}
4006: \varphi_n^R(z)$ are of norm at most $1$ and, for $n\geq 1$, strictly
4007: less than $1$.
4008: 
4009: \item[{\rm{(v)}}] All zeros of $\det(\varphi_n^{R,*}(z))$ and $\det
4010: (\varphi_n^{L,*}(z))$ lie in $\bbC\setminus
4011: \ol{\bbD}$.
4012: 
4013: \item[{\rm{(vi)}}] All zeros of $\det(\varphi_n^R(z))$ and $\det
4014: (\varphi_n^L(z))$ lie in $\bbD$.
4015: \end{SL}
4016: \end{theorem}
4017: 
4018: \begin{remark} (vi) is our first proof of Theorem~\ref{T3.6A}.
4019: \end{remark}
4020: 
4021: \begin{proof} All these results are trivial for $n=0$, so we can hope
4022: to use an inductive argument. So  suppose we have the result for
4023: $n -1$.
4024: 
4025: By \eqref{3.7b},
4026: \begin{equation}\lb{3.24b}
4027: \varphi_n^{R,*} = (\rho_{n-1}^R)^{-1} (\bdone
4028: -z\alpha_{n-1}\varphi_{n-1}^L (\varphi_{n-1}^{R,*})^{-1})
4029: \varphi_{n-1}^{R,*}.
4030: \end{equation}
4031: Since $\abs{\alpha_{n-1}}<1$, if $\abs{z}\leq 1$, each factor on the
4032: right of \eqref{3.24b} is invertible.
4033: This proves (i) and (iii) for $\varphi_n^{R,*}$ and a similar
4034: argument works for $\varphi_n^{L,*}$. If
4035: $z=e^{i\theta}$, $\varphi_n^R (e^{i\theta})=e^{in\theta}\varphi_n^
4036: {R,*} (e^{i\theta})^\dagger$ is also
4037: invertible, so we have (i) and (iii) for $n$.
4038: 
4039: Next, we claim that if $z\in\partial\bbD$, then
4040: \begin{equation}\lb{3.24c}
4041: \varphi_n^{R,*}(z)^\dagger \varphi_n^{R,*}(z) =\varphi_n^L(z)^\dagger
4042: \varphi_n^L(z).
4043: \end{equation}
4044: This follows from taking $z=\xi\in\partial\bbD$ in Proposition~\ref
4045: {P3.6}(a). Given that
4046: $\varphi_n^{R,*}(z)$ is invertible, this implies
4047: \begin{equation}\lb{3.24d}
4048: \bdone =(\varphi_n^L(z) \varphi_n^{R,*}(z)^{-1})^\dagger (\varphi_n^L
4049: (z) \varphi_n^{R,*}(z)^{-1})
4050: \end{equation}
4051: proving the first part of (ii) for $n$. The second part of (ii) is
4052: proven similarly by using Proposition~\ref{P3.6}(b).
4053: 
4054: For $z\in\ol{\bbD}$, let
4055: \[
4056: F(z)=\varphi_n^L(z) \varphi_n^{R,*}(z)^{-1}.
4057: \]
4058: Then $F$ is analytic in $\bbD$, continuous in $\ol{\bbD}$, and $\norm
4059: {F(z)}=1$ on $\partial\bbD$,
4060: so (iv) follows from the maximum principle.
4061: 
4062: Since $\varphi_n^{R,*}(z)$ is invertible on $\ol{\bbD}$, its $\det$
4063: is non-zero there, proving (v).
4064: (vi) then follows from
4065: \begin{equation}\lb{3.24e}
4066: \det(\varphi_n^R(z)) =z^{nl}\, \ol{\det(\varphi_n^{R,*}(1/\bar z))}\, .
4067: \end{equation}
4068: \end{proof}
4069: 
4070: Let $\calV$ be the $\bbC^l$-valued functions on $\partial\bbD$ and
4071: $\calV_n$ the span of the $\bbC^l$-valued polynomials of degree at
4072: most $n$, so
4073: \[
4074: \dim(\calV_n)=\bbC^{l(n+1)}.
4075: \]
4076: Let $\calV_\infty$ be the set $\cup_n \calV_n$ of all
4077: $\bbC^l$-valued polynomials. Let $\pi_n$ be the projection onto
4078: $\calV_n$ in the $\calV$ inner product \eqref {1.33}.
4079: 
4080: It is easy to see that
4081: \begin{equation}\lb{3.24f}
4082: \calV_n \cap \calV_{n-1}^\perp = \{\Phi_n^R(z) v\, \colon v\in\bbC^l\}
4083: \end{equation}
4084: since $\jap{z^l, \Phi_n^R(z) v} =0$ for $l=0,\dots, n-1$ and the
4085: dimensions  on the left and right of \eqref{3.24f} coincide.
4086: $\calV_n \cap \calV_{n-1}^\perp $ can also be described as the set
4087: of $(v^\dagger \Phi_n^L(z))^\dagger$ for $v\in\bbC^l$.
4088: 
4089: We define $M_z \colon \calV_{n-1} \to \calV_n$ or $\calV_\infty
4090: \to \calV_\infty$ as the operator of multiplication by $z$.
4091: 
4092: \begin{theorem}\lb{T3.6D} For all $n$, we have
4093: \begin{equation}\lb{3.24g}
4094: \det_{\bbC^l} (\Phi_n^R(z)) =\det_{\calV_{n-1}} (z\bdone-\pi_{n-1}
4095: M_z \pi_{n-1}).
4096: \end{equation}
4097: \end{theorem}
4098: 
4099: \begin{remarks} 1. Since $\norm{M_z}\leq 1$, \eqref{3.24g}
4100: immediately implies zeros of $\det(\varphi_n^R(z))$
4101: lie in $\ol{\bbD}$, and a small additional argument proves they lie
4102: in $\bbD$. As we will see, this also
4103: implies Theorem~\ref{T3.6B}.
4104: 
4105: \smallskip
4106: 2. Of course, $\pi_{n-1} M_z \pi_{n-1}$ is a cutoff CMV matrix if
4107: written in a CMV basis.
4108: \end{remarks}
4109: 
4110: \begin{proof} If $Q\in \calV_{n-k}$, then by \eqref{3.24f},
4111: \begin{equation}\lb{3.24h}
4112: \pi_{n-1}[(z-z_0)^k Q] = 0 \Leftrightarrow (z-z_0)^k Q=\Phi_n^R(z) v
4113: \end{equation}
4114: for some $v\in\bbC^l$. Thus writing $\det(\Phi_n^R(z)) =\Phi_n^R(z)
4115: v_1 \wedge \dots \wedge
4116: \Phi_n^R(z) v_l$ in a Jordan basis for $\Phi_n^R(z)$, we see that the
4117: order of the zeros of
4118: $\det(\Phi_n^R(z))$ at $z_0$ is exactly the order of $z_0$ as an
4119: algebraic eigenvalue of
4120: $\pi_{n-1} M_z \pi_{n-1}$, that is, the order of $z_0$ as a zero of
4121: the right side of \eqref{3.24g}.
4122: 
4123: Since both sides of \eqref{3.24g} are monic polynomials of degree $nl
4124: $ and their zeros including
4125: multiplicity are the same, we have proven \eqref{3.24g}.
4126: \end{proof}
4127: 
4128: \begin{proof}[Proof of Theorem~\ref{T3.6B}] On the right side of
4129: \eqref{3.24h}, we can put $(\Phi_n^L(z) v^\dagger)^\dagger$ and so
4130: conclude \eqref{3.24g} holds with $\Phi_n^L(z)$ on the left.
4131: 
4132: This proves \eqref{3.24a} if $\varphi$ is replaced by $\Phi$. Since
4133: $\alpha_j^*\alpha_j$ and $\alpha_j\alpha_j^*$ are unitarily equivalent,
4134: $\det(\rho_j^L)=\det (\rho_j^R)$. Thus, $\det(\kappa_n^L) = \det
4135: (\kappa_n^R)$,
4136: and we obtain \eqref{3.24a} for $\varphi$.
4137: \end{proof}
4138: 
4139: It is a basic fact (Theorem~1.7.5 of \cite{S}) that for the scalar
4140: case, any set of $n$ zeros in $\bbD$ are the zeros of a unique
4141: OPUC $\Phi_n$ and any monic polynomial with all its zeros in
4142: $\bbD$ is a monic OPUC. It is easy to see that any set of $nl$
4143: zeros in $\bbD$ is the set of zeros of an OPUC $\Phi_n$, but
4144: certainly not unique. It is an interesting open question to
4145: clarify what matrix monic OPs are monic MOPUCs.
4146: 
4147: \subsection{Bernstein--Szeg\H{o} Approximation} \lb{s3.5A}
4148: 
4149: Given $\{\alpha_j\}_{j=0}^{n-1}\in\bbD^n$, we use Szeg\H{o} recursion
4150: to define polynomials
4151: $\varphi_j^R, \varphi_j^L$ for $j=0,1,\dots, n$. We define a measure
4152: $d\mu_n$ on $\partial\bbD$
4153: by
4154: \begin{equation}\lb{3.32a}
4155: d\mu_n(\theta) = [\varphi_n^R(e^{i\theta})\varphi_n^R (e^{i\theta})^
4156: \dagger]^{-1}
4157: \, \f{d\theta}{2\pi}.
4158: \end{equation}
4159: Notice that \eqref{3.24c} can be rewritten
4160: \begin{equation}\lb{3.32b}
4161: \varphi_n^R(e^{i\theta}) \varphi_n^R(e^{i\theta})^\dagger =
4162: \varphi_n^L (e^{i\theta})^\dagger
4163: \varphi_n^L (e^{i\theta}).
4164: \end{equation}
4165: 
4166: We use here and below the fact that the proof of
4167: Theorem~\ref{T3.6C} only depends on Szeg\H{o} recursion and not on
4168: the a priori existence of a measure. That theorem also shows the
4169: inverse in \eqref{3.32a} exists. Thus,
4170: \begin{equation}\lb{3.32c}
4171: d\mu_n(\theta)= [\varphi_n^L (e^{i\theta})^\dagger \varphi_n^L (e^{i
4172: \theta})]^{-1} \, \f{d\theta}{2\pi}\, .
4173: \end{equation}
4174: 
4175: \begin{theorem} \lb{T3.10A} The measure $d\mu_n$ is normalized
4176: {\rm{(}}i.e., $\mu_n(\partial \bbD)=\bdone${\rm{)}}
4177: and its right MOPUC for $j=0,\dots, n$ are
4178: $\{\varphi_j^R\}_{j=0}^n$, and for $j>n$,
4179: \begin{equation}\lb{3.32d}
4180: \varphi_j^R(z) = z^{j-n} \varphi_n^R(z).
4181: \end{equation}
4182: The Verblunsky coefficients for $d\mu_n$ are
4183: \begin{equation}\lb{3.32e}
4184: \alpha_j (d\mu_n) = \begin{cases} \alpha_j, & j\leq n, \\
4185: \bdzero, & j\geq n+1.
4186: \end{cases}
4187: \end{equation}
4188: \end{theorem}
4189: 
4190: \begin{remarks} 1. In the scalar case, one can multiply by a constant
4191: and renormalize, and then
4192: prove the constant is $1$. Because of commutativity issues, we need a
4193: different argument here.
4194: 
4195: \smallskip
4196: 2. Of course, using \eqref{3.32c}, $\varphi_n^L$ are left MOPUC for $d
4197: \mu_n$.
4198: 
4199: \smallskip
4200: 3. Our proof owes something to the scalar proof in \cite{ENZG91}.
4201: \end{remarks}
4202: 
4203: \begin{proof}
4204: Let $\ang{\cdot,\cdot}_R$ be the inner product associated with $
4205: \mu_n$. By a direct computation, $\ang{\varphi_n^R, \varphi_n^R}_R
4206: =\bdone$, and for $j=0,1,\dots, n-1$,
4207: \begin{align*}
4208: \ang{z^j, \varphi_n^R}_R &= \f{1}{2\pi} \int_0^{2\pi} e^{-ij\theta}
4209: (\varphi_n^R (e^{i\theta})^\dagger)^{-1}\, d\theta \\
4210: &= \f{1}{2\pi i} \oint z^{n-j-1} (\varphi_n^{R,*}(z))^{-1} \, dz =
4211: \bdzero
4212: \end{align*}
4213: by analyticity of $\varphi_n^{R,*}(z)^{-1}$ in $\bbD$ (continuity
4214: in $\ol{\bbD}$).
4215: 
4216: This proves $\varphi_n^R$ is a MOPUC for $d\mu_n$ (and a similar
4217: calculation works for the right side of \eqref{3.32d} if $j\geq
4218: n$). By the inverse Szeg\H{o} recursion and induction downwards,
4219: $\{\varphi_j^R\}_{j=0}^{n-1}$ are also OPs, and by the Szeg\H{o}
4220: recursion, they are normalized. In particular, since
4221: $\varphi_0^R\equiv\bdone$ is normalized, $d\mu_n$ is normalized.
4222: \end{proof}
4223: 
4224: \subsection{Verblunsky's Theorem}\lb{s3.5B}
4225: 
4226: We can now prove the analogue of Favard's theorem for MOPUC; the
4227: scalar version is called Verblunsky's
4228: theorem in \cite{S} after \cite{V35}. A history and other proofs can
4229: be found in \cite{S}. The proof
4230: below is essentially the matrix analogue of that of Geronimus \cite
4231: {Ger46}
4232: (rediscovered in \cite{DGK,ENZG91}).
4233: Delsarte et al. \cite{DGK} presented their proof in the MOPUC case
4234: and they seem to have been the first
4235: with a matrix Verblunsky theorem. One can extend the CMV and the
4236: Geronimus theorem proofs from the scalar
4237: case to get alternate proofs of the theorem below.
4238: 
4239: \begin{theorem}[Verblunsky's Theorem for MOPUC]\lb{T3.10B}
4240: Any sequence $\{\alpha_j\}_{j=0}^\infty \in\bbD^\infty$ is the sequence of
4241: Verblunsky coefficients of a unique measure.
4242: \end{theorem}
4243: 
4244: \begin{proof} Uniqueness is easy, since the $\alpha$'s determine the $
4245: \varphi_j^R$'s and so the
4246: $\Phi_j^R$'s which determine the moments.
4247: 
4248: Given a sequence $\{\alpha_j\}_{j=0}^\infty$, let $d\mu_n$ be the
4249: measures of the last section. By
4250: compactness of $l\times l$ matrix-valued probability measures on
4251: $\partial\bbD$, they have a weak limit. By
4252: using limits, $\{\varphi_j^R\}_{j=0}^\infty$ are the right MOPUC for
4253: $d\mu$ and they determine
4254: the proper Verblunsky coefficients.
4255: \end{proof}
4256: 
4257: \subsection{Matrix POPUC}\lb{s3.5C}
4258: 
4259: Analogously to the scalar case (see \cite
4260: {CMV02,Gol02,JNT89,S308,Wong}), given any unitary $\beta$
4261: in $\calM_l$, we define
4262: \begin{equation}\lb{3.32f}
4263: \varphi_n^R(z;\beta) = z\varphi_{n-1}^R(z) -\varphi_{n-1}^{L,*}(z)
4264: \beta^\dagger.
4265: \end{equation}
4266: As in the scalar case, this is related to the secular determinant of
4267: unitary extensions of the
4268: cutoff CMV matrix. Moreover,
4269: 
4270: \begin{theorem}\lb{T3.10C} Fix $\beta$. All the zeros of $(\varphi_n
4271: (z;\beta))$ lie on $\partial\bbD$.
4272: \end{theorem}
4273: 
4274: \begin{proof} If $\abs{z}<1$, $\varphi_{n-1}^{L,*}(z)$ is invertible and
4275: \[
4276: \varphi_n^R(z;\beta) = -\varphi_{n-1}^{L,*}(z) \beta^\dagger (\bdone -
4277: z\beta \,
4278: \varphi_{n-1}^{L,*}(z)^{-1} \varphi_{n-1}^R(z))
4279: \]
4280: is invertible since the last factor differs from $\bdone$ by a strict
4281: contraction. A similar argument
4282: shows invertibility if $\abs{z}>1$. Thus, the only zeros of $\det
4283: (\dott)$ lie in $\partial\bbD$.
4284: \end{proof}
4285: 
4286: 
4287: 
4288: \subsection{Matrix-Valued Carath\'eodory and Schur Functions}
4289: 
4290: An analytic matrix-valued function $F$ defined on $\D$ is called a
4291: (matrix-valued)
4292: Carath\'eodory function if $F(0) = \bdone$ and $\mathrm{Re} \, F(z)
4293: \equiv \f12 (F(z)+F(z)^\dagger)
4294: \ge 0$ for every $z \in\D$. The following result can be found in \cite[Thm.~2.2.2]{DFK}.
4295: 
4296: \begin{theorem}[Riesz--Herglotz]
4297: If $F$ is a matrix-valued Carath\'eodory function, then there
4298: exists a unique positive semi-definite matrix measure $d\mu$ such
4299: that
4300: \begin{equation}\label{rep}
4301: F(z) = \int \frac{e^{i\theta} + z}{e^{i\theta} - z} \, d\mu(\theta).
4302: \end{equation}
4303: The measure $d\mu$ is given by the unique weak limit of the measures
4304: $d\mu_r(\theta) =
4305: \mathrm{Re} \, F(re^{i\theta}) \f{d\theta}{2\pi}$ as $r \uparrow 1$.
4306: Moreover,
4307: $$
4308: F(z) = c_0 + 2 \sum_{n=1}^\infty c_n z^n
4309: $$
4310: where
4311: $$
4312: c_n = \int e^{-in\theta} \, d\mu(\theta).
4313: $$
4314: Conversely, if $d\mu$ is a positive semi-definite matrix measure,
4315: then \eqref{rep} defines a matrix-valued Carath\'eodory function.
4316: \end{theorem}
4317: 
4318: An analytic matrix-valued function $f$ defined on $\D$ is called a
4319: (matrix-valued) Schur
4320: function if $f(z)^\dagger f(z) \le \bdone$ for every $z \in \D$. This
4321: condition is equivalent
4322: to $f(z) f(z)^\dagger \le \bdone$ for every $z \in \D$ and to $\norm{f
4323: (z)}\leq 1$ for every $z\in
4324: \bbD$. By the maximum principle, if $f$ is not constant, the
4325: inequalities are strict.
4326: The following can be found in \cite[Prop.~4.5.3]{S}:
4327: 
4328: \begin{proposition}
4329: The association
4330: \begin{align}
4331: f(z) & = z^{-1} (F(z) - \bdone)(F(z) + \bdone)^{-1}, \label{Ff}\\
4332: F(z) & = (\bdone + zf(z))(\bdone - zf(z))^{-1} \label{fF}
4333: \end{align}
4334: sets up a one-one correspondence between matrix-valued Carath\'eodory
4335: functions and
4336: matrix-valued Schur functions.
4337: \end{proposition}
4338: 
4339: 
4340: \begin{proposition}\label{prop.3.16}
4341: For $z \in \D$, we have
4342: \begin{equation}\label{ReFfromf}
4343: \mathrm{Re} \, F(z) = (\bdone - \bar z f(z)^\dagger)^{-1} (\bdone - |
4344: z|^2 f(z)^\dagger f(z))
4345: (\bdone - z f(z))^{-1}
4346: \end{equation}
4347: and the non-tangential boundary values $\mathrm{Re} \, F(e^{i\theta})
4348: $ and
4349: $f(e^{i\theta})$ exist for Lebesgue almost every $\theta$.
4350: 
4351: Write $d\mu(\theta) = w(\theta) \frac{d\theta}{2\pi} + d\mu_\s$.
4352: Then, for
4353: almost every $\theta$,
4354: \begin{equation}\label{wandReF}
4355: w(\theta) = \mathrm{Re} \, F(e^{i\theta})
4356: \end{equation}
4357: and for a.e.\ $\theta$, $\det (w(\theta)) \not= 0$ if and only if $f
4358: (e^{i\theta})^\dagger
4359: f(e^{i\theta}) < \bdone$.
4360: \end{proposition}
4361: 
4362: \begin{proof}
4363: The identity \eqref{ReFfromf} follows from \eqref{fF}. The
4364: existence of the boundary values of $f$ follows by application of
4365: the scalar result to the individual entries of $f$. Then
4366: \eqref{fF} gives the boundary values of $F$. We also used the
4367: following fact: Away from a set of zero Lebesgue measure, $\det
4368: (\bdone - z f(z))$ has non-zero boundary values by general
4369: $H^\infty$ theory.
4370: 
4371: \eqref{wandReF} holds for $\jap{\eta, F(z)\eta}_{\C^l}$ and $\jap
4372: {\eta,d\mu\eta}_{\C^l}$ for any $\eta\in\bbC^l$ by the scalar
4373: result. We get \eqref{wandReF} by polarization. From
4374: $$
4375: w(\theta) = (\bdone - e^{-i\theta} f(e^{i\theta})^\dagger)^{-1}
4376: (\bdone - f(e^{i\theta})^\dagger
4377: f(e^{i\theta})) (\bdone - e^{i\theta} f(e^{i\theta}))^{-1}
4378: $$
4379: it follows immediately that $f(e^{i\theta})^\dagger f(e^{i\theta}) <
4380: \bdone$ implies $\det
4381: (w(\theta)) > 0$. Conversely, if $f(e^{i\theta})^\dagger f(e^{i
4382: \theta}) \le \bdone$ but not
4383: $f(e^{i\theta})^\dagger f(e^{i\theta}) < 1$, then $\det (\bdone - f(e^
4384: {i\theta})^\dagger
4385: f(e^{i\theta})) = 0$ and by our earlier arguments $\det (\bdone - e^{-
4386: i\theta}
4387: f(e^{i\theta})^\dagger)^{-1}$ and $\det (\bdone - e^{i\theta} f(e^{i
4388: \theta}))^{-1}$ exist and
4389: are finite; hence $\det (w(\theta)) = 0$. All previous statements are
4390: true away from
4391: suitable sets of zero Lebesgue measure.
4392: \end{proof}
4393: 
4394: 
4395: \subsection{Coefficient Stripping, the Schur Algorithm, and
4396: Geronimus' Theorem} \lb{s3.5} The matrix version of Geronimus'
4397: theorem goes back at least to the book of Bakonyi--Constantinescu
4398: \cite{BakCon}. Let $F(z)$ be the matrix-valued Carath\'eodory
4399: function \eqref{rep} (with the same measure $\mu$ as the one used
4400: in the definition of $ \ang{\cdot,\cdot}_R$). Let us denote
4401: \begin{align*}
4402: u_n^L(z) & = \psi_n^L(z) + \varphi_n^L(z) F(z), \\
4403: u_n^R(z) & = \psi_n^R(z) + F(z) \varphi_n^R(z).
4404: \end{align*}
4405: We also define
4406: \begin{align*}
4407: u_n^{L,*}(z) & = \psi_n^{L,*}(z) - F(z) \varphi_n^{L,*}(z), \\
4408: u_n^{R,*}(z) & = \psi_n^{R,*}(z) - \varphi_n^{R,*}(z) F(z).
4409: \end{align*}
4410: 
4411: 
4412: \begin{proposition}\label{prp8}
4413: For any $\abs{z}<1$, the sequences $u_n^L(z)$, $u_n^R(z)$, $u_n^{L,*}
4414: (z)$, $u_n^{R,*}(z)$
4415: are square summable.
4416: \end{proposition}
4417: \begin{proof}
4418: Denote
4419: $$
4420: f(\theta)= \frac{e^{-i\theta}+\bar z}{e^{-i\theta}-\bar z}\, , \qquad
4421: g(\theta)=\frac{e^{i\theta}+z}{e^{i\theta}-z}\, .
4422: $$
4423: By the definitions \eqref{a1}--\eqref{a3}, we have
4424: \begin{align*}
4425: u_n^L(z)&= \ang{f, \varphi_n^L}_L,
4426: \\
4427: -u_n^{R,*}(z)&= z^n \ang{\varphi_n^R,g}_R,
4428: \\
4429: -u_n^{L,*}(z)&= z^n \ang{\varphi_n^L,g}_L,
4430: \\
4431: u_n^R(z)&= \ang{f, \varphi_n^R}_R.
4432: \end{align*}
4433: Using the Bessel inequality and the fact that $\abs{z}<1$, we obtain
4434: the required
4435: statements.
4436: \end{proof}
4437: 
4438: Next we will consider sequences defined by
4439: \begin{equation}\label{cocy}
4440: \begin{pmatrix} s_n \\ t_n \end{pmatrix} = A^L(\alpha_{n-1},z) \cdots
4441: A^L(\alpha_0,z) \begin{pmatrix} s_0 \\ t_0 \end{pmatrix}
4442: \end{equation}
4443: where $s_n, t_n\in\calM_l$. Similarly, we will consider the sequences
4444: \begin{equation}\label{cocyr}
4445: (s_n,  t_n ) = (s_0, t_0) A^R(\alpha_{0},z) \cdots A^R(\alpha_{n-1},z)
4446: \end{equation}
4447: 
4448: \begin{theorem}\label{th.solu}
4449: Let $z \in \D$ and let  $f$ be the Schur function associated with $d
4450: \mu$ via \eqref{rep}
4451: and \eqref{Ff}. Then:
4452: \begin{SL}
4453: \item[{\rm{(i)}}] A solution of \eqref{cocy} is square summable if
4454: and only if the initial condition is
4455: of the form
4456: $$
4457: \begin{pmatrix} s_0 \\ t_0 \end{pmatrix} = \begin{pmatrix} c \\
4458: z f(z) c \end{pmatrix}
4459: $$
4460: for some matrix $c$ in $\calM_l$.
4461: 
4462: \item[{\rm{(ii)}}] A solution of \eqref{cocyr} is square summable if
4463: and only if the initial condition
4464: is of the form
4465: $$
4466: (  s_0, t_0)  = (c, c z f(z))
4467: $$
4468: for some matrix $c$.
4469: \end{SL}
4470: \end{theorem}
4471: 
4472: \begin{proof}
4473: We shall prove (i); the proof of (ii) is similar.
4474: 
4475: 1. By Proposition~\ref{prp8} and \eqref{*1}, \eqref{*2}, we have
4476: the square summable solution
4477: \begin{equation}
4478: \begin{pmatrix} s_n \\ t_n \end{pmatrix} = \begin{pmatrix} u_n^L(z) \
4479: \ -u_n^{R,*}(z)\end{pmatrix} , \quad
4480: \begin{pmatrix} s_0 \\ t_0 \end{pmatrix} = \begin{pmatrix} d \\ zf(z)
4481: d \end{pmatrix}, \quad
4482: d =(F(z)+\bdone). \label{a9}
4483: \end{equation}
4484: The matrix $d=F(z)+\bdone$ is invertible. Thus, multiplying the above
4485: solution on the right by
4486: $d^{-1}c$ for any given matrix $c$, we get the ``if" part of the
4487: theorem.
4488: 
4489: \smallskip
4490: 2. Let us check that $\varphi_n^{R,*}c$ is not square summable for
4491: any matrix $c\not=0$.
4492: By the CD formula with $\xi=z$, we have
4493: $$
4494: (1-\abs{z}^2) \sum_{k=1}^{n-1} \varphi_k^L(z)^\dagger \varphi_k^L(z) +
4495: \varphi_n^L(z)^\dagger \varphi_n^L(z) = \varphi_n^{R,*}(z)^\dagger
4496: \varphi_n^{R,*}(z)
4497: $$
4498: and so
4499: $$
4500: \varphi_n^{R,*}(z)^\dagger \varphi_n^{R,*}(z) \geq (1-\abs{z}^2)
4501: \varphi_0^L(z)^\dagger \varphi_0^L(z) = (1-\abs{z}^2) \bdone .
4502: $$
4503: Thus, we get
4504: $$
4505: \norm{\varphi_n^{R,*}c}_R2 \ge (1-\abs{z}^2)\Tr c^\dagger c >0.
4506: $$
4507: 
4508: \smallskip
4509: 3. Let $\left(\begin{smallmatrix} s_n  \\ t_n\end{smallmatrix}\right)
4510: $ be any square summable
4511: solution to \eqref{cocy}. Let us write this solution as
4512: \begin{equation}
4513: \begin{pmatrix}s_n \\ t_n \end{pmatrix} = \begin{pmatrix} \varphi_n^L
4514: a \\ \varphi_n^{R,*}a \end{pmatrix}
4515: + \begin{pmatrix} \psi_n^L b \\ -\psi_n^{R,*}b \end{pmatrix},
4516: \quad a= \frac{s_0+t_0}{2}\, , \quad b= \frac{s_0-t_0}{2}\, .
4517: \label{a10}
4518: \end{equation}
4519: Multiplying the solution \eqref{a9} by $b$ and subtracting from \eqref
4520: {a10}, we get a
4521: square summable solution
4522: $$
4523: \begin{pmatrix} \varphi_n^L (z)(a-F(z)b) \\ \varphi_n^{R,*}(a-F(z)b)
4524: \end{pmatrix}.
4525: $$
4526: It follows that $a=F(z)b$, which proves the ``only if" part.
4527: \end{proof}
4528: 
4529: The Schur  function $f$ is naturally associated with $d\mu$ and
4530: hence with the Verblunsky coefficients $\alpha_0, \alpha_1,
4531: \alpha_2, \dots$. The Schur functions obtained by coefficient
4532: stripping will be denoted by $f_1, f_2, f_3, \dots$, that is,
4533: $f_n$ corresponds to Verblunsky coefficients $\alpha_n,
4534: \alpha_{n+1}, \alpha_{n+2}, \dots$. We also write $f_0 \equiv f$.
4535: 
4536: \begin{theorem}[Schur Algorithm and Geronimus' Theorem]\lb{T3.12}
4537: For the Schur functions $f_0, f_1, f_2, \ldots$ associated with
4538: Verblunsky coefficients $\alpha_0, \alpha_1, \alpha_2, \dots$, the
4539: following relations hold:
4540: \begin{align}
4541: f_{n+1}(z) & = z^{-1} (\rho_n^R)^{-1} [f_n(z) - \alpha_n]\,
4542: [\bdone - \alpha_n^\dagger f_n(z)]^{-1} \rho_n^L, \label{salg}\\
4543: f_n(z) & = (\rho_n^R)^{-1} [z f_{n+1}(z) + \alpha_n] \,[\bdone +
4544: z \alpha_n^\dagger f_{n+1}(z)]^{-1} \rho_n^L. \label{salginv}
4545: \end{align}
4546: \end{theorem}
4547: 
4548: \begin{remarks} 1. See \eqref{1.50} and Theorem~\ref{T1.4} to understand
4549: this result.
4550: 
4551: \smallskip
4552: 2. \eqref{1.82a} provides an alternate way to write \eqref{salg} and \eqref{salginv}.
4553: \end{remarks}
4554: 
4555: \begin{proof}
4556: It clearly suffices to consider the case $n = 0$. Consider the
4557: solution of \eqref{cocy}
4558: with initial condition
4559: $$
4560: \begin{pmatrix} \bdone \\ z f_0(z)\end{pmatrix}.
4561: $$
4562: By Theorem~\ref{th.solu}, there exists a matrix $c$ such that
4563: \begin{align*}
4564: \begin{pmatrix} c \\ z f_1(z) c \end{pmatrix} & = A^L(\alpha_0,z)
4565: \begin{pmatrix} \bdone \\ z f_0(z) \end{pmatrix} \\
4566: & = \begin{pmatrix} z (\rho_0^L)^{-1} - z (\rho_0^L)^{-1}
4567: \alpha_0^\dagger f_0(z) \\ - z (\rho_0^R)^{-1} \alpha_0 + z (\rho_0^R)
4568: ^{-1} f_0(z)
4569: \end{pmatrix}.
4570: \end{align*}
4571: {}From this we can compute $zf_1(z)$:
4572: \begin{align*}
4573: z f_1(z) & = [- z (\rho_0^R)^{-1} \alpha_0 + z (\rho_0^R)^{-1} f_0(z)]
4574: \, [z (\rho_0^L)^{-1} - z
4575: (\rho_0^L)^{-1} \alpha_0^\dagger f_0(z)]^{-1} \\
4576: & = (\rho_0^R)^{-1} [f_0(z) - \alpha_0]\, [\bdone -\alpha_0^\dagger
4577: f_0(z)]^{-1} \rho_0^L
4578: \end{align*}
4579: which is \eqref{salg}.
4580: 
4581: Similarly, we can express $f_0$ in terms of $f_1$. From
4582: \begin{align*}
4583: \begin{pmatrix} \bdone \\ z f_0(z) \end{pmatrix}
4584: & = A^L(\alpha_0,z)^{-1} \begin{pmatrix} c \\ z f_1(z) c \end
4585: {pmatrix}  \\
4586: & = \begin{pmatrix} z^{-1} (\rho_0^L)^{-1} & z^{-1} (\rho_0^L)^{-1}
4587: \alpha_0^\dagger \\
4588: (\rho_0^R)^{-1} \alpha_0 & (\rho_0^R)^{-1} \end{pmatrix}
4589: \begin{pmatrix} c \\ z f_1(z) c \end{pmatrix}  \\
4590: & = \begin{pmatrix} z^{-1} (\rho_0^L)^{-1} c +
4591: (\rho_0^L)^{-1} \alpha_0^\dagger f_1(z) c \\ (\rho_0^R)^{-1}
4592: \alpha_0 c + (\rho_0^R)^{-1} z f_1(z) c \end{pmatrix}
4593: \end{align*}
4594: we find that
4595: \begin{align*}
4596: z f_0(z) & = [(\rho_0^R)^{-1} \alpha_0 c + (\rho_0^R)^{-1}
4597: z f_1(z) c] \,[z^{-1} (\rho_0^L)^{-1} c + (\rho_0^L)^{-1} \alpha_0^
4598: \dagger f_1(z) c ]^{-1} \\
4599: & = [(\rho_0^R)^{-1} \alpha_0 + (\rho_0^R)^{-1} z f_1(z)]\, [z^{-1}
4600: (\rho_0^L)^{-1} +
4601: (\rho_0^L)^{-1} \alpha_0^\dagger f_1(z)]^{-1} \\
4602: & = (\rho_0^R)^{-1} [\alpha_0 + z f_1(z)]\, [z^{-1} \bdone + \alpha_0^
4603: \dagger f_1(z)]^{-1} \rho_0^L
4604: \end{align*}
4605: which gives \eqref{salginv}.
4606: \end{proof}
4607: 
4608: 
4609: \subsection{The CMV Matrix}
4610: 
4611: In this section and the next, we discuss CMV matrices for MOPUC.
4612: This was discussed first by Simon in \cite{SimonRev}, which also
4613: has the involved history in the scalar case. Most of the results
4614: in this section appear already in \cite{SimonRev}; the results of
4615: the next section are new here---they parallel the discussion in
4616: \cite[Sect.~4.4]{S} where these results first appeared in the
4617: scalar case.
4618: 
4619: 
4620: \subsubsection{The CMV basis}
4621: 
4622: Consider the two sequences $\chi_n$, $x_n\in \h$, defined by
4623: \begin{alignat*}{2}
4624: \chi_{2k}(z) &= z^{-k} \varphi_{2k}^{L,*}(z) , \qquad & \chi_{2k-1}
4625: (z) &= z^{-k+1}
4626: \varphi_{2k-1}^R(z),
4627: \\
4628: x_{2k}(z) &= z^{-k} \varphi_{2k}^R(z) , \qquad & x_{2k-1}(z) &= z^{-k}
4629: \varphi_{2k-1}^{L,*}(z).
4630: \end{alignat*}
4631: For an integer $k\geq0$, let us introduce the following notation:
4632: $i_k$ is the $(k+1)$th term of the sequence
4633: $0,1,-1,2,-2,3,-3,\dots$, and $j_k$ is the $(k+1)$th term of the
4634: sequence $0,-1,1,-2,2,-3,3,\dots$. Thus, for example, $i_1=1$,
4635: $j_1=-1$.
4636: 
4637: We use the right module structure of $\h$. For a set of functions
4638: $\{f_k(z)\}_{k=0}^n\subset\h$, its module span is the set of all
4639: sums $\sum f_k(z)a_k$ with $a_k\in \calM_l$.
4640: 
4641: \begin{proposition}
4642: \begin{SL}
4643: \item[{\rm{(i)}}] For any $n\geq 1$, the module span of $\{\chi_k\}_
4644: {k=0}^n$ coincides with the module
4645: span of $\{z^{i_k}\}_{k=0}^n$ and the module span of $\{x_k\}_{k=0}^n
4646: $ coincides with the
4647: module span of $\{z^{j_k}\}_{k=0}^n$.
4648: 
4649: \item[{\rm{(ii)}}] The sequences $\{\chi_k\}_{k=0}^\infty$ and $\{x_k
4650: \}_{k=0}^\infty$ are orthonormal:
4651: \begin{equation}
4652: \ang{\chi_k, \chi_m}_R=\ang{x_k, x_m}_R=\delta_{km}. \label{cmv1}
4653: \end{equation}
4654: \end{SL}
4655: \end{proposition}
4656: 
4657: \begin{proof}
4658: (i) Recall that
4659: \begin{align*}
4660: \varphi_n^R(z)&= \kappa_n^R z^n+\lc\{\bdone,\dots, z^{n-1}\},
4661: \\
4662: \varphi_n^{L,*}(z)&= (\kappa_n^L)^\dagger+\lc\{z,\dots, z^{n}\},
4663: \end{align*}
4664: where both $ \kappa_n^R$ and $(\kappa_n^L)^\dagger$ are invertible
4665: matrices. It follows
4666: that
4667: \begin{align*}
4668: \chi_n(z)&=\gamma_n z^{i_n}+\lc\{z^{i_0},\dots, z^{i_{n-1}}\},
4669: \\
4670: x_n(z)&=\delta_n z^{j_n}+\lc\{z^{j_0},\dots, z^{j_{n-1}}\},
4671: \end{align*}
4672: where  $\gamma_n$, $\delta_n$ are invertible matrices. This proves (i).
4673: 
4674: \smallskip
4675: (ii) By the definition of $\varphi_n^L$ and $\varphi_n^R$, we have
4676: \begin{gather}
4677: \ang{\varphi_n^R, \varphi_m^R}_R=\ang{\varphi_n^{L,*}, \varphi_m^
4678: {L,*}}_R =\delta_{nm},
4679: \label{cmv2}
4680: \\
4681: \ang{\varphi_n^R,z^m}_R=\bdzero, \; m=0,\dots,n-1; \quad \ang
4682: {\varphi_n^{L,*},z^m}_R=\bdzero, \;
4683: m=1,\dots,n. \label{cmv3}
4684: \end{gather}
4685: {}From \eqref{cmv2} with $n=m$, we get
4686: $$
4687: \ang{\chi_n,\chi_n}_R=\ang{x_n,x_n}_R=\bdone.
4688: $$
4689: Considering separately the cases of even and odd $n$, it is easy
4690: to prove that
4691: \begin{alignat}{2}
4692: \ang{\chi_n, z^m}_R &=\bdzero, \quad &  m &= i_0,i_1,\dots,i_{n-1},
4693: \label{cmv4}
4694: \\
4695: \ang{x_n, z^m}_R&=\bdzero, \quad & m &= j_0,j_1,\dots,j_{n-1}. \label
4696: {cmv5}
4697: \end{alignat}
4698: For example, for $n=2k$, $m\in\{i_0,\dots, i_{2k-1}\}$ we have
4699: $m+k\in\{1,2,\dots, 2k\}$ and so, by \eqref{cmv3},
4700: $$
4701: \ang{\chi_n,z^m}_R = \ang{z^{-k}\varphi_{2k}^{L,*},z^m}_R =
4702: \ang{\varphi_{2k}^{L,*},z^{m+k}}_R =\bdzero.
4703: $$
4704: The other three cases are considered similarly. From \eqref{cmv4},
4705: \eqref{cmv5}, and (i), we get \eqref{cmv1} for $k\not=m$.
4706: \end{proof}
4707: 
4708: {}From the above proposition and the $\norm{\cdot}_\infty$-density of
4709: Laurent
4710: polynomials in $C(\partial\bbD)$, it follows that  $\{\chi_k\}_{k=0}^
4711: \infty$ and $
4712: \{x_k\}_{k=0}^\infty$ are right orthonormal modula bases in $\h$,
4713: that is,
4714: any element $f\in \h$ can be represented as a
4715: \begin{equation}
4716: f=\sum_{k=0}^\infty \chi_k\ang{\chi_k,f}_R =\sum_{k=0}^\infty x_k\ang
4717: {x_k,f}_R.
4718: \label{cmv6}
4719: \end{equation}
4720: 
4721: 
4722: \subsubsection{The CMV matrix}
4723: 
4724: Consider the matrix of the right homomorphism $f(z) \mapsto zf(z)$
4725: with respect to the
4726: basis $\{\chi_k\}$. Denote ${\mathcal C}_{nm}=\ang{\chi_n, z \chi_m}_
4727: {R}$. The matrix
4728: ${\mathcal C}$ is unitary in the following sense:
4729: $$
4730: \sum_{k=0}^\infty {\mathcal C}_{kn}^\dagger {\mathcal C}_{km} = \sum_
4731: {k=0}^\infty
4732: {\mathcal C}_{nk} {\mathcal C}_{mk}^\dagger=\delta_{nm} \bdone.
4733: $$
4734: The proof follows from \eqref{cmv6}:
4735: \begin{align*}
4736: \delta_{nm}\bdone & = \ang{z\chi_n,z\chi_m}_R = \biggl< \!\!\!
4737: \biggl< \sum_{k=0}^\infty \chi_k
4738: \ang{\chi_k,z\chi_n}_R,z\chi_m \biggr>\!\!\! \biggr>_R =  \sum_{k=0}^
4739: \infty
4740: {\mathcal C}_{kn}^\dagger {\mathcal C}_{km},
4741: \\
4742: \delta_{nm}\bdone &= \ang{\bar z\chi_n,\bar z\chi_m}_R = \biggl< \!\!
4743: \! \biggl< \sum_{k=0}^\infty
4744: \chi_k \ang{\chi_k,\bar z\chi_n}_R, \bar z\chi_m\biggr>\!\!\!
4745: \biggr>_R = \sum_{k=0}^\infty
4746: {\mathcal C}_{nk} {\mathcal C}_{mk}^\dagger.
4747: \end{align*}
4748: 
4749: We note an immediate consequence:
4750: 
4751: \begin{lemma}\label{sol}
4752: Let $\abs{z}\leq1$. Then, for every $m \ge 0$,
4753: $$
4754: \sum_{n = 0}^\infty \chi_n(z) \mathcal{C}_{nm} = z \chi_m(z), \qquad
4755: \sum_{n = 0}^\infty
4756: \mathcal{C}_{mn} \chi_n(1/\bar z)^\dagger = z \chi_m(1/\bar z)^\dagger.
4757: $$
4758: \end{lemma}
4759: 
4760: \begin{proof}
4761: First note that the above series contains only finitely many non-zero
4762: terms. Expanding
4763: $f(z) = z \chi_n$ according to \eqref{cmv6}, we see that
4764: $$
4765: z \chi_n(z) = \sum_{k=0}^\infty \chi_k(z) \ang{\chi_k,z \chi_n}_R =
4766: \sum_{k=0}^\infty
4767: \chi_k(z) \mathcal{C}_{kn}
4768: $$
4769: which is the first identity. Next, taking adjoints, we get
4770: $$
4771: \bar z \chi_n(z)^\dagger = \sum_{k=0}^\infty \mathcal{C}_{kn}^\dagger
4772: \chi_k(z)^\dagger
4773: $$
4774: which yields
4775: \begin{align*}
4776: \bar z \sum_{n = 0}^\infty \mathcal{C}_{mn} \chi_n(z)^\dagger & =
4777: \sum_{n =
4778: 0}^\infty
4779: \mathcal{C}_{mn} \sum_{k=0}^\infty \mathcal{C}_{kn}^\dagger \chi_k(z)^
4780: \dagger \\
4781: & = \sum_{k=0}^\infty \biggl(\, \sum_{n = 0}^\infty \mathcal{C}_{mn} \,
4782: \mathcal{C}_{kn}^\dagger \biggr) \chi_k(z)^\dagger \\
4783: & = \sum_{k=0}^\infty \delta_{mk} \chi_k(z)^\dagger = \chi_m(z)^\dagger.
4784: \end{align*}
4785: Replacing $z$ by $1/\bar z$, we get the required statement.
4786: \end{proof}
4787: 
4788: 
4789: \subsubsection{The $\mathcal{LM}$-representation}
4790: 
4791: Using \eqref{cmv6} for $f=\chi_m$, we obtain:
4792: \begin{equation}
4793: \mathcal{C}_{nm}=\ang{\chi_n, z \chi_m}_{R} = \sum_{k=0}^\infty \ang
4794: {\chi_n,z x_k}_R
4795: \ang{x_k,\chi_m}_R = \sum_{k=0}^\infty \mathcal{L}_{nk} \mathcal{M}_
4796: {km}. \label{cmv7}
4797: \end{equation}
4798: 
4799: Denote by $\Theta(\alpha)$ the $2l \times 2l$ unitary matrix
4800: $$
4801: \Theta(\alpha) = \begin{pmatrix} \alpha^\dagger & \rho^L \\ \rho^R &
4802: - \alpha
4803: \end{pmatrix}.
4804: $$
4805: Using the Szeg\H o recursion formulas \eqref{szright1} and \eqref
4806: {szright4}, we get
4807: \begin{align}
4808: z \varphi_n^R &= \varphi_{n+1}^R\rho_n^R+\varphi_n^{L,*} \alpha_n^
4809: \dagger,
4810: \label{szright5}
4811: \\
4812: \varphi_{n+1}^{L,*} &= \varphi_{n}^{L,*}\rho_n^L-\varphi_{n+1}^{R}
4813: \alpha_n.
4814: \label{szright6}
4815: \end{align}
4816: Taking $n=2k$ and multiplying by $z^{-k}$, we get
4817: \begin{align*}
4818: z x_{2k} &= \chi_{2k} \alpha_{2k}^\dagger+\chi_{2k+1} \rho_{2k}^R,
4819: \\
4820: z x_{2k+1} &= \chi_{2k} \rho_{2k}^L-\chi_{2k+1} \alpha_{2k}.
4821: \end{align*}
4822: It follows that the matrix $\mathcal{L}$ has the structure
4823: $$
4824: \mathcal{L}  = \Theta(\alpha_0) \oplus \Theta(\alpha_2) \oplus \Theta
4825: (\alpha_4) \oplus
4826: \cdots .
4827: $$
4828: Taking $n=2k-1$ in \eqref{szright5}, \eqref{szright6} and multiplying
4829: by $z^{-k}$, we get
4830: \begin{align*}
4831: \chi_{2k-1} &= x_{2k-1} \alpha_{2k-1}^\dagger+x_{2k} \rho_{2k-1}^R,
4832: \\
4833: \chi_{2k} &= x_{2k-1} \rho_{2k-1}^L-x_{2k} \alpha_{2k-1}.
4834: \end{align*}
4835: It follows that the matrix $\mathcal{M}$ has the structure
4836: \begin{equation}
4837: \mathcal{M} = \bdone \oplus \Theta(\alpha_1) \oplus \Theta(\alpha_3)
4838: \oplus \cdots.
4839: \label{*3}
4840: \end{equation}
4841: Substituting this into \eqref{cmv7}, we obtain:
4842: \begin{equation}\label{cmvmatrix}
4843: \mathcal{C} =
4844: \begin{pmatrix} {}& \alpha_0^\dagger & \rho_0^L \alpha_1^\dagger  &
4845: \rho_0^L \rho_1^L & \bdzero & \bdzero & \cdots & {}
4846: \\
4847: {}& \rho_0^R & -\alpha_0 \alpha_1^\dagger & - \alpha_0 \rho_1^L &
4848: \bdzero & \bdzero & \cdots & {}
4849: \\
4850: {}& \bdzero &  \alpha_2^\dagger \rho_1^R & -\alpha_2^\dagger \alpha_1
4851:     & \rho_2^L \alpha_3^\dagger  & \rho_2^L \rho_3^L & \cdots & {}
4852: \\
4853: {}& \bdzero & \rho_2^R \rho_1^R  & -\rho_2^R \alpha_1
4854:     & -\alpha_2 \alpha_3^\dagger & -\alpha_2 \rho_3^L & \cdots & {}
4855: \\
4856: {}& \bdzero & \bdzero & \bdzero & \alpha_4^\dagger \rho_3^R  & -
4857: \alpha_4^\dagger \alpha_3 & \cdots & {}
4858: \\
4859: {}& \vdots & \vdots & \vdots & \vdots & \vdots & \ddots
4860: \end{pmatrix}.
4861: \end{equation}
4862: We note that the analogous formula to this in \cite{SimonRev},
4863: namely, (4.30), is incorrect!
4864: The order of the factors below the diagonal is wrong there.
4865: 
4866: \subsection{The Resolvent of the CMV Matrix}
4867: 
4868: We begin by studying solutions to the equations
4869: \begin{alignat}{2}
4870: \sum_{k=0}^\infty {\mathcal C}_{mk} w_k &= zw_m, \quad & m &\geq 2,
4871: \label{r1}
4872: \\
4873: \sum_{k=0}^\infty \tilde w_k {\mathcal C}_{km} &= z\tilde w_m, \quad
4874: &  m &\geq 1. \label{r2}
4875: \end{alignat}
4876: Let us introduce the following functions:
4877: \begin{align*}
4878: \tilde x_n(z) & =\chi_n(1/\bar{z})^\dagger, \\
4879: \Upsilon_{2n}(z) & =-z^{-n}\psi_{2n}^{L,*}(z), \\
4880: \Upsilon_{2n-1}(z)& = z^{-n+1}\psi_{2n-1}^{R}(z), \\
4881: y_{2n}(z) & = - \Upsilon_{2n}(1/\bar{z})^\dagger =z^{-n}\psi_{2n}^{L}
4882: (z), \\
4883: y_{2n-1}(z) & = -\Upsilon_{2n-1}(1/\bar{z})^\dagger=-z^{-n}\psi_{2n-1}
4884: ^{R,*}(z), \\
4885: p_n(z) & = y_n(z)+\tilde x_n(z)F(z), \\
4886: \pi_n(z) & = \Upsilon_n(z)+F(z)\chi_n(z).
4887: \end{align*}
4888: 
4889: \begin{proposition}\label{sol1}
4890: Let $z\in\D\setminus \{0\}$.
4891: \begin{SL}
4892: \item[{\rm{(i)}}] For each $n\geq 0$, a pair of values $(\tilde w_
4893: {2n}, \tilde w_{2n+1})$
4894: uniquely determines a solution $\tilde w_n$ to \eqref{r2}. Also, for
4895: any pair of values
4896: $(\tilde w_{2n},\tilde w_{2n+1})$  in $\calM_l$, there exists a
4897: solution $\tilde w_{n}$ to
4898: \eqref{r2} with these values at $(2n, 2n+1)$.
4899: 
4900: \item[{\rm{(ii)}}] The set of solutions $\tilde w_n$ to \eqref{r2}
4901: coincides with the set of
4902: sequences
4903: \begin{equation}
4904: \tilde w_n (z)= a \chi_n(z) + b\pi_n(z) \label{r3}
4905: \end{equation}
4906: where $a,b$ range over $\calM_l$.
4907: 
4908: \item[{\rm{(iii)}}] A solution \eqref{r3} is in $\ell^2$ if and only
4909: if $a=0$.
4910: 
4911: \item[{\rm{(iv)}}] A solution \eqref{r3} obeys \eqref{r2} for all $m
4912: \geq0$ if and only if $b=0$.
4913: \end{SL}
4914: \end{proposition}
4915: 
4916: \begin{proposition}\label{sol2}
4917: Let $z\in\D\setminus \{0\}$.
4918: \begin{SL}
4919: \item[{\rm{(i)}}] For each $n\geq 1$, a pair of values $(w_{2n-1},
4920: w_ {2n})$ uniquely determines a solution $w_n$ to \eqref{r1}.
4921: Also, for any pair of values $(w_ {2n-1}, w_{2n})$ in $\calM_l$,
4922: there exists a solution $w_{n}$ to \eqref{r1} with these values at
4923: $(2n-1, 2n)$.
4924: 
4925: \item[{\rm{(ii)}}] The set of solutions $w_n$ to \eqref{r1} coincides
4926: with the set of sequences
4927: \begin{equation}
4928: w_n (z)= \tilde x_n(z)a + p_n(z)b \label{r4}
4929: \end{equation}
4930: where $a,b$ range over $\calM_l$.
4931: 
4932: \item[{\rm{(iii)}}] A solution \eqref{r4} is in $\ell^2$ if and only
4933: if $a=0$.
4934: 
4935: \item[{\rm{(iv)}}] A solution \eqref{r4} obeys \eqref{r1} for all $m
4936: \geq0$ if and only if $b=0$.
4937: \end{SL}
4938: \end{proposition}
4939: 
4940: \begin{proof}[Proof of Proposition~\ref{sol1}]
4941: (i) The matrix ${\mathcal C}-z$ can be written in the form
4942: \begin{equation}
4943: {\mathcal C}-z=
4944: \begin{pmatrix}
4945: A_0 & B_0 & \bdzero & \bdzero & \cdots \\
4946: \bdzero & A_1 & B_1 & \bdzero & \cdots \\
4947: \bdzero & \bdzero & A_2 & B_2 & \cdots \\
4948: \vdots & \vdots & \vdots & \vdots & \ddots\\
4949: \end{pmatrix}
4950: \label{r5}
4951: \end{equation}
4952: where
4953: $$
4954: A_0=
4955: \begin{pmatrix}
4956: \alpha_0^\dagger-z\\ \rho_0^R
4957: \end{pmatrix}
4958: \qquad A_n=
4959: \begin{pmatrix}
4960: \alpha^\dagger_{2n} \rho_{2n-1}^R & -\alpha_{2n}^\dagger \alpha_
4961: {2n-1}-z \\
4962: \rho_{2n}^R\rho_{2n-1}^R & -\rho_{2n}^R \alpha_{2n-1}
4963: \end{pmatrix}
4964: $$
4965: $$
4966: B_n=
4967: \begin{pmatrix}
4968: \rho_{2n}^L\alpha_{2n+1}^\dagger & \rho_{2n}^L \rho_{2n+1}^L \\
4969: -\alpha_{2n}\alpha_{2n+1}^\dagger-z & -\alpha_{2n}\rho_{2n+1}^L
4970: \end{pmatrix}.
4971: $$
4972: Define $\wti W_n=(\tilde w_{2n}, \tilde w_{2n+1})$ for
4973: $n=0,1,2,\dots$. Then \eqref{r2} for $m=2n+1$, $2n+2$ is
4974: equivalent to
4975: $$
4976: \wti W_{n} B_n +\wti W_{n+1} A_{n+1}=\bdzero.
4977: $$
4978: It remains to prove that the $2l\times 2l$ matrices $A_j$, $B_j$
4979: are invertible. Suppose that for some $x,y\in \C^l$, $A_n
4980: \binom{x}{y}=\binom{0}{0}$. This is equivalent to the system
4981: \begin{align*}
4982: \alpha_{2n}^\dagger \rho_{2n-1}^R x - \alpha_{2n}^\dagger \alpha_
4983: {2n-1}y-zy&=\bdzero,
4984: \\
4985: \rho_{2n}^R\rho_{2n-1}^Rx - \rho_{2n}^R\alpha_{2n-1}y&=\bdzero.
4986: \end{align*}
4987: The second equation of this system yields
4988: $\rho_{2n-1}^Rx=\alpha_{2n-1} y$ (since $\rho_{2n}^R$ is
4989: invertible), and upon substitution into the first equation, we get
4990: $y=x=0$. Thus, $\ker(A_n)=\{0\}$. In a similar way, one proves
4991: that $\ker(B_n)=\{0\}$.
4992: 
4993: \smallskip
4994: (ii) First note that $\tilde w_n=\chi_n$ is a solution to \eqref{r2}
4995: by Lemma~\ref{sol}.
4996: Let us check that $\tilde w_n=\Upsilon_n$ is also a solution. If $U_
4997: {km}=(-1)^k
4998: \delta_{km}$, then $(U{\mathcal C}U)_{km}$ for $m\geq1$ coincides
4999: with the CMV matrix
5000: corresponding to the coefficients $\{-\alpha_n\}$. Recall that $
5001: \psi_n^{L,R}$ are the
5002: orthogonal polynomials $\varphi_n^{L,R}$, corresponding to the
5003: coefficients
5004: $\{-\alpha_n\}$. Taking into account the minus signs in the
5005: definition of $\Upsilon_n$,
5006: we see that $\tilde w_n=\Upsilon_n$ solves \eqref{r2} for $m\geq1$.
5007: It follows that any
5008: $\tilde w_n$ of the form \eqref{r3} is a solution to \eqref{r2}.
5009: 
5010: Let us check that any solution to \eqref{r2} can be represented as
5011: \eqref{r3}. By (i), it
5012: suffices to show that for any $\tilde w_0, \tilde w_1$, there exist
5013: $a,b\in\calM_l$ such
5014: that
5015: \begin{align*}
5016: a\chi_0(z)+b\pi_0(z)&=\tilde w_0,
5017: \\
5018: a\chi_1(z)+b\pi_1(z)&=\tilde w_1.
5019: \end{align*}
5020: Recalling that $\chi_0=1$, $\Upsilon_0=-1$,
5021: $\Upsilon_1(z)=(z+\alpha_0^\dagger)(\rho_0^R)^{-1}$,
5022: $\chi_1(z)=(z-\alpha_0^\dagger)(\rho_0^R)^{-1}$, we see that the
5023: above system can be
5024: easily solved for $a,b$ if $z\not=0$.
5025: 
5026: \smallskip
5027: (iii) Let us prove that the solution $\pi_n$ is square integrable. We
5028: will consider
5029: separately the sequences $\pi_{2n}$ and $\pi_{2n-1}$ and prove that
5030: they both belong to
5031: $\ell^2$. By \eqref{a2} and \eqref{a2a}, we have
5032: \begin{align}
5033: \psi_n^R(z)+F(z)\varphi_n^R(z) &= \int
5034: \frac{e^{i\theta}+z}{e^{i\theta}-z} \, d\mu(\theta)\varphi_n^R(e^{i
5035: \theta}), \label{r6}
5036: \\
5037: \psi_n^{L,*}(z)-F(z)\varphi_n^{L,*}(z) &= -z^n\int
5038: \frac{e^{i\theta}+z}{e^{i\theta}-z}\, d\mu(\theta)\varphi_n^L(e^{i
5039: \theta})^\dagger.
5040: \label{r7}
5041: \end{align}
5042: Taking $n=2k$ in \eqref{r7} and $n=2k-1$ in \eqref{r6}, we
5043: get
5044: \begin{align}
5045: \pi_{2k}(z) &= z^k \int  \frac{e^{i\theta}+z}{e^{i\theta}-z}\,
5046: d\mu(\theta)\varphi_{2k}^L (e^{i\theta})^\dagger, \label{r8}
5047: \\
5048: \pi_{2k-1}(z) &=
5049: z^{-k+1} \int  \frac{e^{i\theta}+z}{e^{i\theta}-z} \, d\mu(\theta)
5050: \varphi_{2k-1}^R(e^{i\theta}).
5051: \label{r9}
5052: \end{align}
5053: As $\varphi_{2k}^L$ is an orthonormal sequence, using the Bessel
5054: inequality, from \eqref{r8} we immediately get that $\pi_{2k}$ is
5055: in $\ell^2$.
5056: 
5057: Consider the odd terms $\pi_{2k-1}$. We claim that
5058: \begin{equation}
5059: z^{-k+1} \int
5060: \frac{e^{i\theta}+z}{e^{i\theta}-z}\, d\mu(\theta)\varphi_{2k-1}^R(e^
5061: {i\theta}) = \int
5062: \frac{e^{i\theta}+z}{e^{i\theta}-z}\, d\mu(\theta)e^{i(-k+1)\theta}
5063: \varphi_{2k-1}^R(e^{i\theta}).
5064: \label{r10}
5065: \end{equation}
5066: Indeed, using the right orthogonality of $\varphi_{2k-1}^R$ to $e^{im
5067: \theta}$,
5068: $m=0,1,\dots,2k-2$, we get
5069: \begin{align*}
5070: \int  \frac{e^{i\theta}+z}{e^{i\theta}-z}\, d\mu(\theta) \varphi_
5071: {2k-1}^R (e^{i\theta})
5072: & = \biggl<\!\!\! \biggl< 1+2\sum_{m=1}^\infty \bar z^m e^{im\theta},
5073: \varphi_{2k-1}^R\biggr>\!\!\! \biggr>_R \\
5074: & = \biggl<\!\!\! \biggl<  2 \!\!\! \sum_{m=2k-1}^\infty \bar z^m
5075: e^{im\theta},\varphi_{2k-1}^R\biggr>\!\!\! \biggr>_R
5076: \end{align*}
5077: and
5078: \begin{align*}
5079: \int \frac{e^{i\theta}+z}{e^{i\theta}-z}z^{k-1}\, &e^{i(-k+1)\theta} \,
5080: d\mu(\theta)\varphi_{2k-1}^R(e^{i\theta}) = \\
5081: & = \biggl<\!\!\! \biggl< \bar z^{k-1}e^{i(k-1)\theta} \biggl(1+2\sum_
5082: {m=1}^\infty \bar z^m
5083: e^{im\theta}\biggr),\varphi_{2k-1}^R\biggr>\!\!\! \biggr>_R \\
5084: & = \biggl<\!\!\! \biggl< 2 \!\!\! \sum_{m=2k-1}^\infty \bar z^m
5085: e^{im\theta},\varphi_{2k-1}^R\biggr>\!\!\! \biggr>_R
5086: \end{align*}
5087: which proves \eqref{r10}. The identities \eqref{r10} and \eqref{r9}
5088: yield
5089: $$
5090: \pi_{2k-1}(z) = \biggl<\!\!\! \biggl<
5091: \frac{e^{-i\theta}+\bar z}{e^{-i\theta}-\bar z},\chi_{2k-1}\biggr>\!\!
5092: \! \biggr>_R
5093: $$
5094: and, since $\chi_{2k-1}$ is a right orthogonal sequence, the Bessel
5095: inequality ensures
5096: that $\pi_{2k-1}(z)$ is in $\ell^2$. Thus, $\pi_k(z)$ is in $\ell^2$.
5097: 
5098: Next, as in the proof  of Theorem~\ref{th.solu}, using the CD
5099: formula, we check that the sequence $\norm{\varphi_n^{L,*}(z)}_R$
5100: is bounded below and therefore the sequence $\chi_{2n}(z)$ is not
5101: in $\ell^2$. This proves the statement (iii).
5102: 
5103: \smallskip
5104: (iv) By Lemma~\ref{sol}, the solution $\chi_n(z)$ obeys \eqref{r2}
5105: for all $m\geq0$. It is easy to check directly that the solution
5106: $\pi_n(z)$ does not obey \eqref{r2} for $m=0$ if $z\not=0$. This
5107: proves the required statement.
5108: \end{proof}
5109: 
5110: \begin{proof}[Proof of Proposition~\ref{sol2}]
5111: (i) For $j=1,2,\dots$, define $W_j=(w_{2j-1}, w_{2j})$. Then, using
5112: the block structure
5113: \eqref{r5}, we can rewrite \eqref{r1} for $m=2j, 2j+1$ as $A_j  W_j
5114: +B_{j}W_{j+1}=\bdzero$. By
5115: the proof of Proposition~\ref{sol1}, the matrices $A_j$ and $B_j$ are
5116: invertible, which
5117: proves (i).
5118: 
5119: \smallskip
5120: (ii) Lemma~\ref{sol} ensures that $\tilde x_n(z)$ is a solution of
5121: \eqref{r1}. As in the
5122: proof of Proposition~\ref{sol1}, by considering the matrix $(U
5123: {\mathcal C}U)_{km}$, one
5124: checks that $y_n(z)$ is also a solution to \eqref{r1}.
5125: 
5126: Let us prove that any solution to \eqref{r1} can be represented in
5127: the form \eqref{r4}.
5128: By (i), it suffices to show that for any $w_1$, $w_2$, there exist
5129: $a,b\in\calM_l$ such that
5130: \begin{align*}
5131: \tilde x_1(z) a + p_1(z) b & = w_1,
5132: \\
5133: \tilde x_2(z) a + p_2(z) b & = w_2.
5134: \end{align*}
5135: We claim that this system of equations can be solved for $a$, $b$.
5136: Here are the main
5137: steps. Substituting the definitions of $\tilde x_n(z)$ and $p_n(z)$,
5138: we rewrite this
5139: system as
5140: \begin{align*}
5141: \varphi_1^{R,*} (a+F(z)b) - \psi_1^{R,*}(z) b & = z w_1,
5142: \\
5143: \varphi_2^{L} (a+F(z)b) + \psi_2^{L}(z) b & = z w_2.
5144: \end{align*}
5145: Using Szeg\H o recurrence, we can substitute the expressions for
5146: $\varphi_2^{L}$, $\psi_2^{L}$, which helps rewrite our system as
5147: \begin{align*}
5148: \varphi_1^{R,*} (a+F(z)b) - \psi_1^{R,*}(z) b & = z w_1,
5149: \\
5150: \varphi_1^{L} (a+F(z)b) + \psi_1^{L}(z) b & = \rho_1^L
5151: w_2+\alpha_1^\dagger  w_1.
5152: \end{align*}
5153: Substituting explicit formulas for $\varphi_1^{R,*}$,
5154: $\varphi_1^{L}$, $ \psi_1^{R,*}$, $\psi_1^{L}$, and expressing
5155: $F(z)$ in terms of $f(z)$, we can rewrite this as
5156: \begin{align*}
5157: (\rho_0^R)^{-1}(1-\alpha_0z)a+2z(\rho_0^R)^{-1}(f(z)-\alpha_0)(\bdone-
5158: zf(z))^{-1}b
5159: & = z w_1,
5160: \\
5161: (\rho_0^L)^{-1}(z-\alpha_0^\dagger)a+2z(\rho_0^L)^{-1}(\bdone-
5162: \alpha_0^\dagger
5163: f(z))(\bdone-zf(z))^{-1}b & = \rho_1^L w_2+\alpha_1^\dagger  w_1.
5164: \end{align*}
5165: Denote
5166: \begin{align*}
5167: a_1 &=(\rho_0^R)^{-1}(1-\alpha_0z)a,
5168: \\
5169: b_1 &=2z(\rho_0^L)^{-1}(\bdone-\alpha_0^\dagger
5170: f(z))(\bdone-zf(z))^{-1}b.
5171: \end{align*}
5172: Then in terms of $a_1$, $b_1$, our system can be rewritten as
5173: \begin{align*}
5174: a_1+X_1b_1 & = z w_1,
5175: \\
5176: X_2a_1+b_1 & = \rho_1^L w_2+\alpha_1^\dagger  w_1,
5177: \end{align*}
5178: where
5179: \begin{align*}
5180: X_1 &=(\rho_0^R)^{-1}(f(z)-\alpha_0)(\bdone-\alpha_0
5181: f(z))^{-1}\rho_0^L,
5182: \\
5183: X_2 &=(\rho_0^L)^{-1}(z-\alpha_0^\dagger)(\bdone-\alpha_0
5184: z)^{-1}\rho_0^R.
5185: \end{align*}
5186: Since $\norm{f(z)}<1$ and $\abs{z}<1$, we can apply
5187: Corollary~\ref{C1.3.5}, which yields $\norm{X_1}<1$ and
5188: $\norm{X_2}<1$. It follows that our system can be solved for
5189: $a_1$, $b_1$.
5190: 
5191: \smallskip
5192: (iii) As $p_n(z)=-\pi_n(1/\bar z)^\dagger$, by Proposition~\ref
5193: {sol1}, we get that
5194: $p_n(z)$ is in $\ell^2$. In the same way, as $\tilde
5195: x_n(z)=\chi_n(1/\bar z)^\dagger$, we get that $\tilde x_n(z)$ is not
5196: in $\ell^2$.
5197: 
5198: \smallskip
5199: (iv) By Lemma~\ref{sol}, the solution $\tilde x_n(z)$ obeys \eqref
5200: {r1} for all $m\geq0$.
5201: Using the explicit formula for $y_n(z)$, one easily checks that the
5202: solution $y_n(z)$
5203: does not obey \eqref{r1} for $m=0,1$.
5204: \end{proof}
5205: 
5206: \begin{theorem} \lb{T3.21}
5207: We have for $z \in \D$,
5208: $$
5209: [(\mathcal{C} - z)^{-1}]_{k,l} =
5210: \begin{cases} (2z)^{-1} \tilde x_k(z) \pi_l(z), & l > k \text{
5211: or } k = l \text{ even}, \\
5212: (2z)^{-1} p_k(z) \chi_l(z), & k > l \text{ or } k = l \text{
5213: odd}.
5214: \end{cases}
5215: $$
5216: \end{theorem}
5217: 
5218: \begin{proof}
5219: Fix $z \in \D$. Write $G_{k,l}(z) = [(\mathcal{C} -
5220: z)^{-1}]_{k,l}$. Then $G_{\bddot , l}(z)$ is equal to
5221: $(\mathcal{C} - z)^{-1} \delta_{l}$, which means that $G_{k,l}(z)$
5222: solves \eqref{r1} for $m \not= l$. Since $G_{\bddot , l}(z)$ is
5223: $\ell^2$ at infinity and obeys the equation at $m = 0$, we see
5224: that it is a right-multiple of $p$ for large $k$ and a
5225: right-multiple of $\tilde x$ for small $k$. Thus,
5226: $$
5227: G_{k,l}(z) =
5228: \begin{cases} \tilde x_k(z) a_l(z), & k < l \text{ or } k = l  \text
5229: { even}, \\
5230: p_k(z) b_l(z), & k > l \text{ or } k = l \text{ odd}.
5231: \end{cases}
5232: $$
5233: Similarly,
5234: $$
5235: G_{k,l}(z) = \begin{cases} \tilde b_k(z) \pi_l(z), & k < l \text
5236: { or } k = l \text{ even}, \\
5237: \tilde a_k(z) \chi_l(z), & k > l \text{ or } k = l \text{ odd}.
5238: \end{cases}
5239: $$
5240: Equating the two expressions, we find
5241: \begin{alignat}{2}
5242: \label{rfp1} \tilde x_k(z) a_l(z) &= \tilde b_k(z) \pi_l(z)  \quad &
5243: k &< l \text{ or } k = l \text{ even}, \\
5244: \label{rfp2} p_k(z) b_l(z) &= \tilde a_k(z) \chi_l(z)  \quad& k &> l
5245: \text{ or } k = l
5246: \text{ odd}.
5247: \end{alignat}
5248: Putting $k = 0$ in \eqref{rfp1} and setting $\tilde b_0(z) = c_1(z)$,
5249: we find $a_l(z) =
5250: c_1(z) \pi_l(z)$. Putting $l = 0$ in \eqref{rfp2} and setting $b_0(z)
5251: = c_2(z)$, we find
5252: $p_k(z) c_2(z) = \tilde a_k(z)$. Thus,
5253: $$
5254: G_{k,l}(z) =
5255: \begin{cases} \tilde x_k(z) c_1(z) \pi_l(z), & k < l \text{ or } k =
5256: l  \text{ even}, \\
5257: p_k(z) c_2(z) \chi_l(z), & k > l \text{ or } k = l \text{ odd}.
5258: \end{cases}
5259: $$
5260: We claim that $c_1(z) = c_2(z) = (2z)^{-1} \bdone$. Consider the
5261: case $k = l = 0$. Then, on the one hand, by the definition,
5262: \begin{align}
5263: G_{0,0}(z) & = \int \frac{1}{e^{i\theta} - z} \, d\mu (e^{i\theta})
5264: \notag
5265: \\
5266: & = \int (2z)^{-1} \biggl[ \frac{e^{i\theta} + z}{e^{i\theta} - z} -
5267: 1 \biggr] d\mu (e^{i\theta}) \notag
5268: \\
5269: &= (2z)^{-1} (F(z) - \bdone)
5270: \label{spthm}
5271: \end{align}
5272: and on the other hand,
5273: $$
5274: G_{0,0}(z) = \tilde x_0(z) c_1(z) \pi_0(z) = c_1(z) (F(z) - \bdone).
5275: $$
5276: This shows $c_1(z) = (2z)^{-1} \bdone$. Next, consider the case $k=1
5277: $, $l=0$. Then, on the one
5278: hand, by the definition,
5279: $$
5280: G_{1,0}(z) = \ang{\chi_1, (e^{i\theta}-z)^{-1}\chi_0}_{R}
5281: $$
5282: and on the other hand,
5283: $$
5284: G_{1,0}(z) = p_1(z) c_2(z) \chi_0(z).
5285: $$
5286: Let us calculate the expressions on the right-hand side. We have
5287: \begin{equation}
5288: p_1(z) c_2(z) \chi_0(z) = (\rho_0^R)^{-1}(-z^{-1}-\alpha_0+(z^{-1}-
5289: \alpha_0)F(z))c_2(z)
5290: \label{r11}
5291: \end{equation}
5292: and
5293: \begin{align*}
5294: \ang{\chi_1, &(e^{i\theta}-z)^{-1}\chi_0}_{R} = \\
5295: & = (\rho_0^R)^{-1} \int (e^{-i\theta}-\alpha_0)d\mu(\theta)(e^{i
5296: \theta}-z)^{-1}
5297: \\
5298: & = (\rho_0^R)^{-1}\int [z^{-1}(e^{i\theta}-z)^{-1}-z^{-1}e^{-i
5299: \theta}-\alpha_0
5300: (e^{i\theta}-z)^{-1}] \, d\mu(\theta)
5301: \\
5302: & = (\rho_0^R)^{-1}\biggl[\frac{1}{2z2}\, (F(z)-\bdone)-\frac{1}{2z}
5303: \, \alpha_0 (F(z)-\bdone) -
5304: \frac{1}{z}\int e^{-i\theta}\, d\mu(\theta)\biggr].
5305: \end{align*}
5306: Taking into account the identity
5307: $$
5308: \int e^{-i\theta}d\mu(\theta)=\alpha_0
5309: $$
5310: (which can be obtained, e.g., by expanding $\ang{\varphi_1^R,
5311: \varphi_0^R}_R=0$),
5312: we get
5313: $$
5314: \ang{\chi_1, (e^{i\theta}-z)^{-1}\chi_0}_{R} = \frac{1}{2z}\,
5315: (\rho_0^R)^{-1}
5316: (-z^{-1}-\alpha_0+(z^{-1}-\alpha_0)F(z)).
5317: $$
5318: Comparing this with \eqref{r11}, we get $c_2(z)=(2z)^{-1}\bdone$.
5319: \end{proof}
5320: 
5321: As an immediate corollary, evaluating the kernel on the diagonal for
5322: even and odd
5323: indices, we obtain the formulas
5324: \begin{align}
5325: \int \varphi_{2n}^L(e^{i\theta})\, \frac{d\mu(\theta)}{e^{i\theta}-z}\,
5326: \varphi_{2n}^L(e^{i\theta})^\dagger &= -\frac1{2 z^{2n+1}}\, \varphi_
5327: {2n}^L(z)
5328: u_{2n}^{L,*}(z), \label{r12}
5329: \\
5330: \int \varphi_{2n-1}^R(e^{i\theta})^\dagger\, \frac{d\mu(\theta)}{e^{i
5331: \theta}-z}\,
5332: \varphi_{2n-1}^R(e^{i\theta}) & = -\frac1{2 z^{2n}}\, u_{2n-1}^{R,*}(z)
5333: \varphi_{2n-1}^R(z). \label{r13}
5334: \end{align}
5335: Combining this with \eqref{wronsk1} and \eqref{wronsk2}, we find
5336: \begin{align}
5337: u_n^L(z) \varphi_n^{L,*}(z) + \varphi_n^L(z) u_n^{L,*}(z) & = 2 z^n,
5338: \label{wronsk5} \\
5339: \varphi_n^{R,*}(z) u_n^R(z) + u_n^{R,*}(z) \varphi_n^R(z) & = 2 z^n.
5340: \label{wronsk6}
5341: \end{align}
5342: 
5343: 
5344: \subsection{Khrushchev Theory} \lb{s3.13}
5345: 
5346: Among the deepest and most elegant methods in OPUC are those of
5347: Khrushchev
5348: \cite{Kh2000,Khr,KhGo}. We have not been able to extend them to
5349: MOPUC! We regard
5350: their extension as an important open question; we present the first
5351: very partial
5352: steps here.
5353: 
5354: Let
5355: $$
5356: \Omega = \{ \theta \,\colon \det w(\theta) > 0 \}.
5357: $$
5358: 
5359: \begin{theorem}
5360: For every $n \ge 0$,
5361: $$
5362: \{ \theta : f_n(e^{i\theta})^\dagger f_n(e^{i\theta}) < \bdone \} =
5363: \Omega
5364: $$
5365: up to a set of zero Lebesgue measure.
5366: 
5367: Consequently,
5368: \begin{equation}\label{fnintlower}
5369: \int \| f_n (e^{i\theta}) \| \, \frac{d\theta}{2\pi} \ge 1 - \frac{|
5370: \Omega|}{2\pi}\, .
5371: \end{equation}
5372: \end{theorem}
5373: 
5374: \begin{proof}
5375: Recall that, by Proposition~\ref{prop.3.16}, up to a set of zero
5376: Lebesgue measure,
5377: $$
5378: \{ \theta\, \colon f_0(e^{i\theta})^\dagger f_0(e^{i\theta}) < \bdone
5379: \} =
5380: \{ \theta\, \colon \det w(\theta) > 0 \}
5381: $$
5382: so, by induction, it suffices to show that, up to a set of zero
5383: Lebesgue measure,
5384: $$
5385: \{ \theta : f_0(e^{i\theta})^\dagger f_0(e^{i\theta}) < \bdone \}
5386: = \{ \theta : f_1(e^{i\theta})^\dagger f_1(e^{i\theta}) < \bdone
5387: \}.
5388: $$
5389: This in turn follows from the fact that the Schur algorithm, which
5390: relates the two functions, preserves the property $g^\dagger g <
5391: \bdone$.
5392: 
5393: Notice that away from $\Omega$, $f_n (e^{i\theta})$ has norm one and
5394: therefore,
5395: $$
5396: \int \| f_n (e^{i\theta}) \| \, \frac{d\theta}{2\pi} \ge \int_
5397: {\Omega^c} \| f_n
5398: (e^{i\theta}) \| \, \frac{d\theta}{2\pi} =  \int_{\Omega^c} 1 \, \frac
5399: {d\theta}{2\pi}
5400: $$
5401: which yields \eqref{fnintlower}.
5402: \end{proof}
5403: 
5404: Define
5405: $$
5406: b_n(z;d\mu) = \varphi_n^L(z;d\mu) \varphi_n^{R,*}(z;d\mu)^{-1}.
5407: $$
5408: 
5409: \begin{proposition}
5410: \begin{SL}
5411: \item[{\rm{(a)}}] $b_{n+1} = (\rho_n^L)^{-1} (z b_n - \alpha_n^
5412: \dagger) (\bdone - z
5413: \alpha_n b_n)^{-1} \rho_n^R$.
5414: 
5415: \item[{\rm{(b)}}] The Verblunsky coefficients of $b_n$ are $(-
5416: \alpha_{n-1}^ \dagger, - \alpha_{n-2}^\dagger, \dots , -
5417: \alpha_0^\dagger, \bdone)$.
5418: \end{SL}
5419: \end{proposition}
5420: 
5421: \begin{proof}
5422: (a) By the Szeg\H{o} recursion, we have that
5423: \begin{align*}
5424: b_{n+1} & = \varphi_{n+1}^L (\varphi_{n+1}^{R,*})^{-1} \\
5425: & = ((\rho_n^L)^{-1} z \varphi_n^L - (\rho_n^L)^{-1}
5426: \alpha_n^\dagger \varphi_n^{R,*}) ((\rho_n^R)^{-1}
5427: \varphi_n^{R,*} - z (\rho_n^R)^{-1} \alpha_n \varphi_n^L)^{-1} \\
5428: & = (\rho_n^L)^{-1} (z \varphi_n^L - \alpha_n^\dagger \varphi_n^{R,*})
5429: (\varphi_n^{R,*} - z \alpha_n \varphi_n^L)^{-1} \rho_n^R \\
5430: & = (\rho_n^L)^{-1} (z \varphi_n^L (\varphi_n^{R,*})^{-1} -
5431: \alpha_n^\dagger \varphi_n^{R,*} (\varphi_n^{R,*})^{-1})  \\
5432: & \qquad \qquad \qquad (\varphi_n^{R,*} (\varphi_n^{R,*})^{-1} - z
5433: \alpha_n \varphi_n^L
5434: (\varphi_n^{R,*})^{-1})^{-1} \rho_n^R \\
5435: & = (\rho_n^L)^{-1} ( z b_n - \alpha_n^\dagger) (\bdone - z
5436: \alpha_n b_n )^{-1} \rho_n^R.
5437: \end{align*}
5438: 
5439: \smallskip
5440: (b) It follows from part (a) that the first Verblunsky coefficient
5441: of $b_n$ is $- \alpha_{n-1}^\dagger$ and that its first Schur
5442: iterate is $b_{n-1}$; compare Theorem~\ref{T3.12}. This gives the
5443: claim by induction and the fact that $b_0 = \bdone$.
5444: \end{proof}
5445: 
5446: 
5447: \section{The Szeg\H{o} Mapping and the Geronimus Relations} \lb{s4}
5448: 
5449: In this chapter, we present the matrix analogue of the Szeg\H{o}
5450: mapping and the resulting Geronimus relations. This establishes a
5451: correspondence between certain matrix-valued measures on the unit
5452: circle and matrix-valued measures on the interval $[-2,2]$ and,
5453: consequently, a correspondence between Verblunsky coefficients and
5454: Jacobi parameters. Throughout this chapter, we will denote
5455: measures on the circle by $d\mu_C$ and measures on the interval by
5456: $d\mu_I$.
5457: 
5458: The scalar versions of these objects are due to Szeg\H{o} \cite
5459: {Sz22a} and Geronimus \cite{Ger46}. There are four proofs that we
5460: know of: the original argument of Geronimus \cite{Ger46} based on
5461: Szeg\H{o}'s formula in \cite{Sz22a}, a proof of Damanik--Killip
5462: \cite{DKacta} using Schur functions, a proof of Killip--Nenciu
5463: \cite{KN} using CMV matrices, and a proof of Faybusovich--Gekhtman
5464: \cite{FG99} using canonical moments.
5465: 
5466: The matrix version of these objects was studied by
5467: Yakhlef--Marcell\'an \cite{YM} who proved Theorem~\ref{T4.2} below
5468: using the Geronimus--Szeg\H{o} approach. Our proof uses the
5469: Killip--Nenciu--CMV approach. In comparing our formula with \cite
5470: {YM}, one needs the following dictionary (their objects on the
5471: left of the equal sign and ours on the right):
5472: \begin{align*}
5473: H_n &= -\alpha_{n+1}^\dagger ,\\
5474: D_n &= A_n, \\
5475: E_n &= B_{n+1}.
5476: \end{align*}
5477: 
5478: Dette--Studden \cite{DS} have extended the theory of canonical
5479: moments from OPRL to MOPRL.
5480: It would be illuminating to use this to extend the proof that
5481: Faybusovich--Gekhtman
5482: \cite{FG99} gave of Geronimus relations for scalar OPUC to MOPUC.
5483: 
5484: Suppose $d\mu_C$ is a non-trivial positive semi-definite Hermitian
5485: matrix measure on the unit circle that is invariant under $\theta
5486: \mapsto - \theta$ (i.e., $z \mapsto \bar z =z^{-1}$). Then we
5487: define the measure $d\mu_I$ on the interval $[-2,2]$ by
5488: $$
5489: \int f(x) \, d\mu_I(x) = \int f(2 \cos \theta) \, d\mu_C(\theta)
5490: $$
5491: for $f$ measurable on $[-2,2]$. The map
5492: $$
5493: \mathrm{Sz} : d\mu_C \mapsto d\mu_I
5494: $$
5495: is called the Szeg\H{o} mapping.
5496: 
5497: The Szeg\H{o} mapping can be inverted as follows. Suppose $d\mu_I$
5498: is a non-degenerate positive semi-definite matrix measure on
5499: $[-2,2]$. Then we define the measure $d\mu_C$ on the unit circle
5500: which is invariant under $\theta \mapsto - \theta$ by
5501: $$
5502: \int g(\theta) \, d\mu_C(\theta) = \int g \left( \arccos (x/2)
5503: \right) \, d\mu_I(x)
5504: $$
5505: for $g$ measurable on $\partial \D$ with $g(\theta) = g(-\theta)$.
5506: 
5507: We first show that for the measures on the circle of interest in this
5508: section, the
5509: Verblunsky coefficients are always Hermitian.
5510: 
5511: \begin{lemma}\label{alphaherm}
5512: Suppose $d\mu_C$ is a non-trivial positive semi-definite
5513: Hermitian matrix measure on the unit circle. Denote the associated
5514: Verblunsky coefficients by $ \{ \alpha_n \}$. Then, $d\mu_C$ is
5515: invariant under $\theta \mapsto - \theta$ if and only if
5516: $\alpha_n^ \dagger = \alpha_n$ for every $n$.
5517: \end{lemma}
5518: 
5519: \begin{proof}
5520: For a polynomial $P$, denote $\tilde P(z)=P(\bar z)^\dagger$.
5521: 
5522: 1. Suppose that $d\mu_C$ is invariant under $\theta \mapsto - \theta$.
5523: Then we have
5524: $$
5525: \ang{f,g}_L=\ang{\tilde g,\tilde f}_R
5526: $$
5527: for all $f$, $g$. Inspecting the orthogonality conditions which
5528: define $\Phi_n^L$ and $\Phi_n^R$, we see that
5529: \begin{equation}
5530: \tilde\Phi_n^L=\Phi^R_n \text{ and }
5531: \ang{\Phi_n^L,\Phi_n^L}_L=\ang{\Phi_n^R,\Phi_n^R}_R. \label{4.1}
5532: \end{equation}
5533: Next, we claim that
5534: \begin{equation}
5535: \kappa_n^L=\kappa_n^{R,\dagger}.
5536: \label{4.2}
5537: \end{equation}
5538: Indeed, recall the definition of $\kappa_n^L$, $\kappa_n^R$:
5539: \begin{alignat*}{2}
5540: \kappa_n^L&=u_n\ang{\Phi_n^L,\Phi_n^L}_L^{-1/2},
5541: \text{ $u_n$ is unitary, } \kappa_{n+1}^L(\kappa_n^L)^{-1}>0, \quad &
5542: \kappa_0^L &=1,
5543: \\
5544: \kappa_n^R&=\ang{\Phi_n^R,\Phi_n^R}_R^{-1/2}v_n,
5545: \text{ $v_n$ is unitary, } (\kappa_n^R)^{-1}\kappa_{n+1}^R>0, \quad &
5546: \kappa_0^R &=1.
5547: \end{alignat*}
5548: Using this definition and \eqref{4.1}, one can easily prove by induction
5549: that $v_n=u_n^\dagger$ and therefore \eqref{4.2} holds true.
5550: 
5551: Next, taking $z=0$ in \eqref{szmonic}, we get
5552: \begin{align*}
5553: \alpha_n=-(\kappa_n^R)^{-1}\Phi_{n+1}^L(0)^\dagger(\kappa_n^L)^\dagger,
5554: \\
5555: \alpha_n=-(\kappa_n^R)^\dagger\Phi_{n+1}^R(0)^\dagger (\kappa_n^L)^{-1}.
5556: \end{align*}
5557: {}From here and \eqref{4.1}, \eqref{4.2}, we get $\alpha_n=\alpha_n^
5558: \dagger$.
5559: 
5560: \smallskip
5561: 2. Assume $\alpha_n^\dagger=\alpha_n$ for all $n$. Then, by
5562: Theorem~\ref{szethm}(c), we have $\rho_n^L=\rho_n^R$. It follows
5563: that in this case the Szeg\H{o} recurrence relation is invariant
5564: with respect to the change $\varphi_n^L\mapsto \tilde
5565: \varphi_n^R$, $\varphi_n^R\mapsto \tilde \varphi_n^L$. It follows
5566: that $\varphi_n^L= \tilde \varphi_n^R$, $\varphi_n^R=\tilde
5567: \varphi_n^L$. In particular, we get
5568: \begin{equation}
5569: \ang{\varphi_n^L,\varphi_m^L}_L=\ang{\tilde\varphi_m^R, \tilde
5570: \varphi_n^R}_R.
5571: \label{4*}
5572: \end{equation}
5573: Now let $f$ and $g$ be any polynomials; we have
5574: $$
5575: f(z)=\sum_n f_n \varphi^L_n(z),
5576: \quad
5577: \tilde f(z)=\sum_n \tilde \varphi_n^L(z) f_n^\dagger,
5578: $$
5579: and a similar expansion for $g$. Using these expansions and \eqref
5580: {4*}, we get
5581: $$
5582: \ang{f,g}_L=\ang{\tilde g,\tilde f}_R \text{ for all polynomials $f$,
5583: $g$.}
5584: $$
5585: {}From here it follows that the measure $d\mu_C$ is invariant
5586: under $\theta \mapsto - \theta$.
5587: \end{proof}
5588: 
5589: Now consider two measures $d\mu_C$ and $d\mu_I =
5590: \mathrm{Sz}(d\mu_C)$ and the associated CMV and Jacobi matrices.
5591: What are the relations between the parameters of these matrices?
5592: 
5593: \begin{theorem} \lb{T4.2}
5594: Given $d\mu_C$ and $d\mu_I = \mathrm{Sz}(d\mu_C)$ as above, the
5595: coefficients of the associated CMV and Jacobi matrices satisfy the
5596: Geronimus relations:
5597: \begin{align}
5598: B_{k+1} & = \sqrt{\bdone - \alpha_{2k-1}} \; \alpha_{2k}
5599: \sqrt{\bdone-\alpha_{2k-1}} -
5600: \sqrt{\bdone+\alpha_{2k-1}} \; \alpha_{2k-2} \sqrt{\bdone +\alpha_
5601: {2k-1}}\, ,  \lb{4.3}\\
5602: A_{k+1} & = \sqrt{\bdone - \alpha_{2k-1}} \sqrt{\bdone-\alpha_{2k}^2}
5603: \sqrt{\bdone+\alpha_{2k+1}}\, . \lb{4.4}
5604: \end{align}
5605: \end{theorem}
5606: 
5607: \begin{remarks} 1. For these formulas to hold for $k=0$,
5608: we set $\alpha_{-1} = - {\boldsymbol 1}$.
5609: 
5610: \smallskip
5611: 2. There are several proofs of the Geronimus relations in the
5612: scalar case. We follow the proof given by Killip and Nenciu in \cite
5613: {KN}.
5614: 
5615: \smallskip
5616: 3. These $A$'s are, in general, not type~1 or 2 or 3.
5617: \end{remarks}
5618: 
5619: \begin{proof}
5620: For a Hermitian $l \times l$ matrix $\alpha$ with $\| \alpha \| < 1$,
5621: define the unitary
5622: $2l \times 2l$ matrix $S(\alpha)$ by
5623: $$
5624: S(\alpha) = \frac{1}{\sqrt{2}} \begin{pmatrix} \sqrt{\bdone - \alpha}
5625: & - \sqrt{\bdone + \alpha} \\
5626: \sqrt{\bdone + \alpha} & \sqrt{\bdone - \alpha} \end{pmatrix}.
5627: $$
5628: Since $\alpha^\dagger = \alpha$, the associated $\rho^L$ and $\rho^R$
5629: coincide and will
5630: be denoted by $\rho$. We therefore have
5631: $$
5632: \Theta(\alpha) = \begin{pmatrix} \alpha & \rho \\ \rho & - \alpha
5633: \end{pmatrix}
5634: $$
5635: and hence, by a straightforward calculation,
5636: \begin{align*}
5637: S(\alpha) \Theta(\alpha) S(\alpha)^{-1} & = \frac{1}{2} \begin
5638: {pmatrix} \sqrt{\bdone - \alpha} & - \sqrt{\bdone + \alpha} \\
5639: \sqrt{\bdone + \alpha} & \sqrt{\bdone - \alpha} \end{pmatrix}  \begin
5640: {pmatrix} \alpha & \rho \\
5641: \rho & - \alpha
5642: \end{pmatrix}  \begin{pmatrix} \sqrt{\bdone - \alpha} & \sqrt{\bdone
5643: + \alpha} \\
5644: - \sqrt{\bdone + \alpha} & \sqrt{\bdone - \alpha} \end{pmatrix} \\
5645: & = \begin{pmatrix} - {\boldsymbol 1} & {\boldsymbol 0} \\
5646: {\boldsymbol 0} & {\boldsymbol
5647: 1} \end{pmatrix}.
5648: \end{align*}
5649: 
5650: Thus, if we define
5651: $$
5652: \mathcal{S} = \bdone\oplus S(\alpha_1)\oplus S(\alpha_3)\oplus \dots
5653: $$
5654: and
5655: $$
5656: \mathcal{R} = \bdone\oplus(-\bdone)\oplus \bdone\oplus(-\bdone)\oplus
5657: \dots,
5658: $$
5659: it follows that (see \eqref{*3})
5660: $$
5661: \mathcal{S} \mathcal{M} \mathcal{S}^\dagger = \mathcal{R}.
5662: $$
5663: 
5664: The matrix $\mathcal{L} \mathcal{M} + \mathcal{M} \mathcal{L}$ is
5665: unitarily equivalent to
5666: $$
5667: \mathcal{A} = \mathcal{S} ( \mathcal{L} \mathcal{M} +
5668: \mathcal{M} \mathcal{L} ) \mathcal{S}^\dagger
5669: =
5670: \mathcal{S}
5671: \mathcal{L} \mathcal{S}^\dagger \mathcal{R} + \mathcal{R}
5672: \mathcal{S} \mathcal{L} \mathcal{S}^\dagger.
5673: $$
5674: Observe that $\mathcal{A}$ is the direct sum of two block Jacobi
5675: matrices. Indeed, it follows quickly from the explicit form of
5676: $\mathcal{R}$ that the even-odd and odd-even block entries of
5677: $\mathcal{A}$ vanish. Consequently, $\mathcal{A}$ is the direct sum
5678: of its odd-odd and even-even block entries. We will call these two
5679: block Jacobi matrices $J$ and $\tilde J$, respectively.
5680: 
5681: Consider $\mathcal C$ as an operator on $\h_v$.
5682: Then $d\mu_C$ is the spectral measure of $\mathcal{C}$
5683: in the following sense:
5684: $$
5685: [\mathcal{C} ^m]_{0,0}=\int e^{im\theta} d\mu_C(\theta);
5686: $$
5687: see \eqref{spthm}. Then, by the spectral theorem, $d\mu_C(-\theta)$
5688: is the spectral measure of $\mathcal{C}^{-1}$ and so
5689: $d\mu_I$ is the spectral measure of
5690: $\mathcal{C}+\mathcal{C}^{-1}=
5691: \mathcal{L} \mathcal{M} + (\mathcal{L} \mathcal{M})^{-1}=
5692: \mathcal{L} \mathcal{M} +\mathcal{M} \mathcal{L}$. Since  $\mathcal{S}
5693: $ leaves $\begin{pmatrix}
5694: {\boldsymbol 1} & {\boldsymbol 0} & {\boldsymbol 0} & {\boldsymbol
5695: 0} & \cdots
5696: \end{pmatrix}^t$ invariant, we see that $d\mu_I$ is the spectral
5697: measure of $J$.
5698: 
5699: To determine the block entries of $J$, we only need to compute the
5700: odd-odd block entries of $\mathcal{A}$. For $k \ge 0$, we have
5701: \begin{align*}
5702: \mathcal{A}_{2k+1,2k+1} & = \begin{pmatrix} \sqrt{\bdone+\alpha_
5703: {2k-1}} &
5704: \sqrt{\bdone - \alpha_{2k-1}} \end{pmatrix}  \begin{pmatrix} -
5705: \alpha_{2k-2} & {\boldsymbol 0} \\ {\boldsymbol 0} & \alpha_{2k}
5706: \end{pmatrix}
5707: \begin{pmatrix} \sqrt{\bdone+\alpha_{2k-1}} \\ \sqrt{\bdone-\alpha_
5708: {2k-1}}
5709: \end{pmatrix} \\
5710: & = \sqrt{\bdone - \alpha_{2k-1}} \; \alpha_{2k} \sqrt{\bdone-\alpha_
5711: {2k-1}} -
5712: \sqrt{\bdone+\alpha_{2k-1}} \; \alpha_{2k-2} \sqrt{\bdone+\alpha_{2k-1}}
5713: \end{align*}
5714: and
5715: \begin{align*}
5716: \mathcal{A}_{2k+1,2k+3} & = \begin{pmatrix} \sqrt{\bdone+\alpha_
5717: {2k-1}} &
5718: \sqrt{\bdone - \alpha_{2k-1}} \end{pmatrix}  \begin{pmatrix}
5719: {\boldsymbol
5720: 0} & {\boldsymbol 0} \\ \rho_{2k} & {\boldsymbol 0}
5721: \end{pmatrix}
5722: \begin{pmatrix} \sqrt{\bdone+\alpha_{2k+1}} \\ \sqrt{\bdone-\alpha_{2k
5723: +1}}
5724: \end{pmatrix} \\
5725: & = \sqrt{\bdone - \alpha_{2k-1}} \sqrt{\bdone-\alpha_{2k}^2}
5726: \sqrt{\bdone+\alpha_{2k+1}}\, .
5727: \end{align*}
5728: The result follows.
5729: \end{proof}
5730: 
5731: As in \cite{KN,S2}, one can also use this to get Geronimus relations
5732: for the
5733: second Szeg\H{o} map.
5734: 
5735: \section{Regular MOPRL} \lb{s5}
5736: 
5737: The theory of regular (scalar) OPs was developed by Stahl--Totik \cite
5738: {StT} generalizing
5739: a definition of Ullman \cite{Ull} for $[-1,1]$. (See Simon \cite
5740: {EqMC} for a review and
5741: more about the history.) Here we develop the basics for MOPRL; it is
5742: not hard to do the same
5743: for MOPUC.
5744: 
5745: \subsection{Upper Bound and Definition} \lb{s5.1}
5746: 
5747: \begin{theorem}\lb{T5.1} Let $d\mu$ be a non-trivial $l\times l$
5748: matrix-valued measure
5749: on $\bbR$ with $E=\supp(d\mu)$ compact. Then {\rm{(}}with $C(E)=$
5750: logarithmic capacity
5751: of $E${\rm{)}}
5752: \begin{equation} \lb{5.1}
5753: \limsup_{n\to\infty}\, \abs{\det(A_1\cdots A_n)}^{1/n} \leq C(E)^l.
5754: \end{equation}
5755: \end{theorem}
5756: 
5757: \begin{remarks} 1. $\abs{\det(A_1\cdots A_n)}$ is constant over
5758: equivalent Jacobi parameters.
5759: 
5760: \smallskip
5761: 2. For the scalar case, this is a result of Widom \cite{Wid} (it
5762: might be older) whose proof
5763: extends to the matrix case.
5764: \end{remarks}
5765: 
5766: \begin{proof} Let $T_n$ be the Chebyshev polynomials for $E$ (see
5767: \cite[Appendix~B]{EqMC} for a
5768: definition) and let $T_n^{(l)}$ be $T_n\otimes\bdone$, that is, the $l
5769: \times l$ matrix polynomial
5770: obtained by multiplying $T_n(x)$ by $\bdone$. $T_n^{(l)}$ is monic
5771: so, by \eqref{2.7d} and \eqref{2.32a},
5772: \begin{align*}
5773: \abs{\det(A_1 \cdots A_n)}^{1/n}
5774: & \leq \biggl| \det\biggl( \int \abs{T_n^{(l)}(x)}^2\, d\mu(x)\biggr)
5775: \biggr|^{1/2n} \\
5776: &\leq \sup_n\, \abs{T_n(x)}^{l/n}.
5777: \end{align*}
5778: By a theorem of Szeg\H{o} \cite{Sz24}, $\sup_n \abs{T_n(x)}^{1/n}\to C
5779: (E)$, so \eqref{5.1} follows.
5780: \end{proof}
5781: 
5782: \begin{definition} Let $d\mu$ be a non-trivial $l\times l$ matrix-valued measure with $E=\supp(d\mu)$
5783: compact. We say $\mu$ is {\it regular\/} if equality holds in \eqref
5784: {5.1}.
5785: \end{definition}
5786: 
5787: \subsection{Density of Zeros} \lb{s5.2}
5788: 
5789: The following is a simple extension of the scalar results
5790: (see \cite{DLS} or \cite[Sect.~2]{EqMC}):
5791: 
5792: \begin{theorem}\lb{T5.2} Let $d\mu$ be a regular measure with $E=\supp
5793: (d\mu)$. Let $d\nu_n$
5794: be the zero counting measure for $\det(p_n^L(x))$, that is, if $\{x_j^
5795: {(n)}\}_{j=1}^{nl}$ are
5796: the zeros of this determinant {\rm{(}}counting degenerate zeros
5797: multiply{\rm{)}}, then
5798: \begin{equation} \lb{5.2}
5799: d\nu_n = \f{1}{nl} \sum_{j=1}^{nl} \delta_{x_j^{(n)}}.
5800: \end{equation}
5801: Then $d\nu_n$ converges weakly to $d\rho_E$, the equilibrium measure
5802: for $E$.
5803: \end{theorem}
5804: 
5805: \begin{remark} For a discussion of capacity, equilibrium measure,
5806: quasi-every (q.e.), etc.,
5807: see \cite[Appendix~A]{EqMC} and \cite{Helms,Land,Ran}.
5808: \end{remark}
5809: 
5810: \begin{proof} By \eqref{2.54L} and \eqref{2.54x},
5811: \begin{equation} \lb{5.3}
5812: \abs{\det(p_n^R(x))} \geq \bigl(\tfrac{d}{D}\bigr)^l
5813: \bigl(1+\bigl(\tfrac{d}{D}\bigr)^2\bigr)^{(n-1)l/2}
5814: \end{equation}
5815: so, in particular,
5816: \begin{equation} \lb{5.4}
5817: \liminf_n\, \abs{\det(p_n^R(x))}^{1/nl} \geq
5818: \bigl(1+\bigl(\tfrac{d}{D}\bigr)^2 \bigr)^{1/2} \geq 1.
5819: \end{equation}
5820: But
5821: \begin{equation} \lb{5.5}
5822: \abs{\det(p_n^R(x))} = \abs{\det(A_1\cdots A_n)}^{-1} \exp (-nl\Phi_
5823: {\nu_n}(x))
5824: \end{equation}
5825: where $\Phi_\nu$ is the potential of the measure $\nu$. Let $\nu_
5826: \infty$ be a limit point of
5827: $\nu_n$ and use equality in \eqref{5.1} and \eqref{5.4} to conclude,
5828: for $x\notin\cvh(E)$,
5829: \[
5830: \exp(-\Phi_{\nu_\infty}(x))\geq C(E)
5831: \]
5832: which, as in the proof of Theorem~2.4 of \cite{EqMC}, implies that $
5833: \nu_\infty=\rho_e$.
5834: \end{proof}
5835: 
5836: The analogue of the almost converse of this last theorem has an extra
5837: subtlety relative to
5838: the scalar case:
5839: 
5840: \begin{theorem}\lb{T5.3} Let $d\mu$ be a non-trivial $l\times l$
5841: matrix-valued measure on
5842: $\bbR$ with $E=\supp(d\mu)$ compact. If $d\nu_n\to d\rho_E$, then
5843: either $\mu$ is regular
5844: or else, with $d\mu=M(x)\, d\mu_\tr(x)$, there is a set $S$ of
5845: capacity zero, so $\det(M(x))=0$
5846: for $d\mu_\tr$-a.e.\ $x\notin S$.
5847: \end{theorem}
5848: 
5849: \begin{remark} By taking direct sums of regular point measures, it is
5850: easy to find regular
5851: measures where $\det(M(x))=0$ for $d\mu_\tr$-a.e.\ $x$.
5852: \end{remark}
5853: 
5854: \begin{proof} For a.e.\ $x$ with $\det(M(x))\neq 0$, we have (see
5855: Lemma~\ref{L5.5} below)
5856: \[
5857: p_n^R(x)\leq C(n+1)\bdone.
5858: \]
5859: The theorem then follows from the proof of Theorem~2.5 of \cite{EqMC}.
5860: \end{proof}
5861: 
5862: \subsection{General Asymptotics} \lb{s5.3}
5863: 
5864: The following generalizes Theorem~1.10 of \cite{EqMC} from OPRL to
5865: MOPRL---it is the matrix analogue
5866: of a basic result of Stahl--Totik \cite{StT}. Its proof is
5867: essentially the same as in \cite{EqMC}. By $\sigma_\ess(\mu)$, we
5868: mean the
5869: essential spectrum of the block Jacobi matrix associated to $\mu$.
5870: 
5871: \begin{theorem}\lb{T5.4} Let $E\subset\bbR$ be compact and let $\mu$
5872: be an $l\times l$ matrix-valued
5873: measure of compact support with $\sigma_\ess(\mu)=E$.
5874: Then the following are equivalent:
5875: \begin{SL}
5876: \item[{\rm{(i)}}] $\mu$ is regular, that is, $\lim_{n\to\infty} \abs
5877: {\det(A_1 \cdots A_n)}^{1/n}
5878: =C(E)^l$.
5879: 
5880: \item[{\rm{(ii)}}] For all $z$ in $\bbC$, uniformly on compacts,
5881: \begin{equation}\lb{5.6}
5882: \limsup \abs{\det(p_n^R(z))}^{1/n} \leq e^{G_E(z)}.
5883: \end{equation}
5884: 
5885: \item[{\rm{(iii)}}] For q.e.\ $z$ in $E$, we have
5886: \begin{equation}\lb{5.7}
5887: \limsup \abs{\det(p_n^R(z))}^{1/n}\leq 1.
5888: \end{equation}
5889: \end{SL}
5890: 
5891: \smallskip
5892: Moreover, if {\rm{(i)--(iii)}} hold, then
5893: \begin{SL}
5894: \item[{\rm{(iv)}}] For every $z\in\bbC\setminus\cvh(\supp(d\mu))$, we
5895: have
5896: \begin{equation}\lb{5.8}
5897: \lim_{n\to\infty}\, \abs{\det(p_n^R(z))}^{1/n} = e^{G_E(z)}.
5898: \end{equation}
5899: 
5900: \item[{\rm{(v)}}] For q.e.\ $z\in\partial\Omega$,
5901: \begin{equation}\lb{5.9}
5902: \limsup_{n\to\infty} \, \abs{\det(p_n^R(z))}^{1/n} =1.
5903: \end{equation}
5904: \end{SL}
5905: \end{theorem}
5906: 
5907: \begin{remarks} 1. $G_E$, the potential theorists' Green's function
5908: for $E$, is defined by $G_E(z)=
5909: -\log(C(E))-\Phi_{\rho_E}(z)$.
5910: 
5911: \smallskip
5912: 2. There is missing here one condition from Theorem~1.10 of \cite
5913: {EqMC} involving general polynomials.
5914: Since the determinant of a sum can be much larger than the sum of the
5915: determinants, it is not obvious
5916: how to extend this result.
5917: \end{remarks}
5918: 
5919: \subsection{Weak Convergence of the CD Kernel and Consequences} \lb
5920: {s5.3A}
5921: 
5922: The results of Simon in \cite{Weak-cd} extend to the matrix case. The
5923: basic result is:
5924: 
5925: \begin{theorem}\lb{T5.4A} The measures $d\nu_n$ and $\f{1}{(n+1)l}\Tr
5926: (K_n(x,x))\, d\mu(x)$
5927: have the same weak limits. In particular, if $d\mu$ is regular,
5928: \begin{equation}\lb{5.9a}
5929: \f{1}{(n+1)l}\, \Tr (K_n(x,x)d\mu(x)) \overset{w}{\longrightarrow} d
5930: \rho_E.
5931: \end{equation}
5932: \end{theorem}
5933: 
5934: As in \cite{Weak-cd}, $(\pi_n M_x \pi_n)^j$ and $(\pi_n M_x^j \pi_n)$
5935: have a difference of
5936: traces which is bounded as $n\to\infty$, and this implies the result.
5937: Once one has this,
5938: combining it with Theorem~\ref{T2.18B} leads to:
5939: 
5940: \begin{theorem}\lb{T5.4B} Let $I=[a,b]\subset E\subset\bbR$ with $E$
5941: compact. Let $\sigma_\ess
5942: (d\mu)=E$ for an $l\times l$ matrix-valued measure, and suppose $d\mu$
5943: is regular for $E$ and
5944: \begin{equation}\lb{5.9b}
5945: d\mu = W(x)\, dx + d\mu_\s
5946: \end{equation}
5947: where $d\mu_\s$ is singular and $\det(W(x))>0$ for a.e.\ $x\in I$. Then,
5948: \begin{alignat}{2}
5949: & \text{\rm{(1)}} \qquad && \lim_{n\to\infty} \int_I p_n^R(x)^\dagger
5950: \, d\mu_\s (x)
5951: p_n^R(x)\to \boldsymbol{0}, \lb{5.9c} \\
5952: & \text{\rm{(2)}} \qquad && \int_I\, \biggl\| \f{1}{n+1} \sum_{j=0}^n
5953: p_j^R(x)^\dagger
5954: W(x) p_j(x) -\rho_E(x)\bdone \biggr\|\, dx \to 0. \lb{5.9d}
5955: \end{alignat}
5956: \end{theorem}
5957: 
5958: 
5959: \subsection{Widom's Theorem} \lb{s5.4}
5960: 
5961: \begin{lemma}\lb{L5.5} Let $d\mu$ be an $l\times l$ matrix-valued
5962: measure supported on a compact
5963: $E\subset\bbR$ and let $d\eta$ be a scalar measure on $\bbR$ so
5964: \begin{equation} \lb{5.10}
5965: d\mu(x) =W(x)\, d\eta(x) + d\mu_\s (x)
5966: \end{equation}
5967: where $d\mu_\s$ is $d\eta$ singular. Suppose for $d\eta$-a.e.\ $x$,
5968: \begin{equation} \lb{5.11}
5969: \det(W(x)) >0.
5970: \end{equation}
5971: Then for $d\eta$-a.e.\ $x$, there is a positive real function $C(x)$
5972: so that
5973: \begin{equation} \lb{5.12}
5974: \norm{p_n^R(x)} \leq C(x) (n+1)\bdone.
5975: \end{equation}
5976: In particular,
5977: \begin{equation} \lb{5.13}
5978: \abs{\det(p_n^R(x))} \leq C(x)^l (n+1)^l.
5979: \end{equation}
5980: \end{lemma}
5981: 
5982: \begin{proof} Since $\norm{p_n^R}_R^2=1$, we have
5983: \begin{equation} \lb{5.14}
5984: \sum_{n=0}^\infty (n+1)^{-2} \norm{p_n^R}_R^2 <\infty
5985: \end{equation}
5986: so
5987: \[
5988: \sum_{n=0}^\infty (n+1)^{-2} \Tr (p_n^R(x)^\dagger W(x) p_n^R(x)) <
5989: \infty
5990: \]
5991: for $d\eta$-a.e.\ $x$. Since \eqref{5.11} holds, for a.e.\ $x$,
5992: \[
5993: W(x)\geq b(x)\bdone
5994: \]
5995: for some scalar function $b(x)$. Thus, for a.e.\ $x$,
5996: \[
5997: \sum_{n=0}^\infty (n+1)^{-2} \Tr(p_n^R(x)^\dagger p_n^R(x)) \leq C(x)^2.
5998: \]
5999: 
6000: Since $\norm{A}^2\leq \Tr(A^\dagger A)$, we find \eqref{5.12}, which
6001: in turn implies \eqref{5.13}.
6002: \end{proof}
6003: 
6004: This lemma replaces Lemma~4.1 of \cite{EqMC} and then the proof there
6005: of Theorem~1.12 extends to give
6006: (a matrix version of the theorem of Widom \cite{Wid}):
6007: 
6008: \begin{theorem}\lb{T5.6} Let $d\mu$ be an $l\times l$ matrix-valued
6009: measure with $\sigma_\ess (d\mu)
6010: =E\subset\bbR$ compact. Suppose
6011: \begin{equation} \lb{5.15}
6012: d\mu(x) =W(x)\, d\rho_E(x) + d\mu_\s (x)
6013: \end{equation}
6014: with $d\mu_\s$ singular with respect to $d\rho_E$. Suppose for $d
6015: \rho_E$-a.e.\ $x$, $\det(W(x))>0$.
6016: Then $\mu$ is regular.
6017: \end{theorem}
6018: 
6019: 
6020: \subsection{A Conjecture} \lb{s5.5}
6021: 
6022: We end our discussion of regular MOPRL with a conjecture---an analog
6023: of a theorem of Stahl--Totik
6024: \cite{StT}; see also Theorem~1.13 of \cite{EqMC} for a proof and
6025: references. We expect the key will be
6026: some kind of matrix Remez inequality. For direct sums, this
6027: conjecture follows from Theorem~1.13 of
6028: \cite{EqMC}.
6029: 
6030: \begin{conjecture}\lb{Con5.7} Let $E$ be a finite union of disjoint
6031: closed intervals in $\bbR$. Suppose
6032: $\mu$ is an $l\times l$ matrix-valued measure on $\bbR$ with $\sigma_
6033: \ess (d\mu) =E$. For each $\eta>0$
6034: and $m=1,2,\dots$, define
6035: \begin{equation} \lb{5.16}
6036: S_{m,\eta} = \{x \,\colon \mu([x-\tfrac{1}{m}, x+\tfrac{1}{m}])\geq e^
6037: {-\eta m}\bdone\}.
6038: \end{equation}
6039: Suppose that for each $\eta$ (with $\abs{\cdot}=$ Lebesgue measure)
6040: \[
6041: \lim_{m\to\infty} \, \abs{E\setminus S_{m,\eta}} =0.
6042: \]
6043: Then $\mu$ is regular.
6044: \end{conjecture}
6045: 
6046: \medskip
6047: 
6048: \noindent{\bf Acknowledgments.} It is a pleasure to thank Alexander Aptekarev, Christian Berg,
6049: Antonio Dur\'an, Jeff Geronimo, Fritz~Gesztesy, Alberto Gr\"unbaum, Paco Marcell\'an,
6050: Ken McLaughlin, Hermann Schulz-Baldes, and Walter Van Assche for useful correspondence.
6051: D.D.\ was supported in part by NSF grants DMS--0500910 and
6052: DMS-0653720. B.S.\ was supported in part by NSF grants DMS--0140592
6053: and DMS-0652919 and U.S.--Israel Binational Science Foundation
6054: (BSF) Grant No.\ 2002068.
6055: 
6056: 
6057: \bigskip
6058: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6059: \begin{thebibliography}{888}
6060: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
6061: 
6062: %
6063: \bibitem{ALK} R.~Ackner, H.~Lev-Ari, and T.~Kailath,
6064: \emph{The Schur algorithm for matrix-valued meromorphic functions},
6065: SIAM J.\ Matrix Anal.\ Appl. \textbf{15} (1994), 140--150.
6066: %
6067: \bibitem{AG} N.~Akhiezer and I.~Glazman,
6068: \textit{Theory of Linear Operators in Hilbert Space},
6069: Dover Publications, New York (1993).
6070: %
6071: \bibitem{Alp} D.~Alpay,
6072: \textit{The Schur Algorithm, Reproducing Kernel Spaces and System Theory},
6073: SMF/AMS Texts and Monographs \textbf{5}, American Mathematical
6074: Society, Providence, R.I.; Soci\'et\'e Math\'ematique de France, Paris (2001).
6075: %
6076: \bibitem{AN} A.~Aptekarev and E.~Nikishin,
6077: \emph{The scattering problem for a discrete Sturm--Liouville operator},
6078: Mat.\ Sb. \textbf{121(163)} (1983), 327--358.
6079: %
6080: \bibitem{Atk} F.~V.~Atkinson,
6081: \textit{Discrete and Continuous Boundary Problems},
6082: Mathematics in Science and Engineering \textbf{8},
6083: Academic Press, New York-London (1964).
6084: %
6085: \bibitem{BakCon} M.~Bakonyi and T.~Constantinescu,
6086: \textit{Schur's Algorithm and Several Applications},
6087: Pitman Research Notes in Math. \textbf{261},
6088: Longman, Essex, U.K. (1992).
6089: %
6090: \bibitem{BB83} S.~Basu and N.~K.~Bose,
6091: \emph{Matrix Stieltjes series and network models},
6092: SIAM J.\ Math.\ Anal. \textbf{14} (1983), 209--222.
6093: %
6094: \bibitem{BGMS03} E.~D.~Belokolos, F.~Gesztesy, K.~A.~Makarov,
6095: and L.~A.~Sakhnovich,
6096: \emph{Matrix-valued generalizations of the theorems of Borg and Hochstadt},
6097: Evolution Equations (G.~Ruiz Goldstein et al., eds.), pp.~1--34,
6098: Lecture Notes in Pure and Applied Mathematics \textbf{234},
6099: Marcel Dekker, New York (2003).
6100: 
6101: %
6102: \bibitem{Ber} Ju.~M.~Berezans'ki,
6103: \textit{Expansions in Eigenfunctions of Selfadjoint Operators},
6104: Translations of Mathematical Monographs \textbf{17},
6105: American Mathematical Society, R.I. (1968).
6106: %
6107: \bibitem{Berg} C.~Berg,
6108: \emph{The matrix moment problem},
6109: to appear in Coimbra Lecture Notes on Orthogonal Polynomials
6110: (A.~Foulqui\'e Moreno and Amilcar Branquinho, eds.).
6111: %
6112: \bibitem{Berg-Duran1} C.~Berg and A.~J.~Dur{\'a}n,
6113: \emph{Measures with finite index of determinacy or a mathetmatical model
6114: for Dr.~Jekyll and Mr.~Hyde},
6115: Proc.\ Amer.\ Math.\ Soc. \textbf{125} (1997), 523--530.
6116: %
6117: \bibitem{Berg-Duran2} C.~Berg and A.~J.~Dur{\'a}n,
6118: \emph{Orthogonal polynomials and analytic functions associated to positive definite matrices},
6119: J.\ Math.\ Anal.\ Appl. \textbf{315} (2006), 54--67.
6120: %
6121: \bibitem{Bha} R.~Bhatia,
6122: \textit{Matrix Analysis},
6123: Graduate Texts in Mathematics \textbf{169}, Springer-Verlag, New York (1997).
6124: %
6125: \bibitem{CFMV} M.~J.~Cantero, M.~P.~Ferrer, L.~Moral, and L.~Vel\'azquez,
6126: \emph{A connection between orthogonal polynomials on the unit circle
6127: and matrix orthogonal polynomials on the real line},
6128: J.\ Comput.\ Appl.\ Math. \textbf{154} (2003), 247--272.
6129: %
6130: \bibitem{CMV02} M.~J.~Cantero, L.~Moral, and L.~Vel\'azquez,
6131: \emph{Measures and para-orthogonal polynomials on the unit circle},
6132: East J.\ Approx.\ {\bf 8} (2002),  447--464.
6133: %
6134: \bibitem{CMV05} M.~J.~Cantero, L.~Moral, and L.~Vel\'azquez,
6135: \emph{Differential properties of matrix orthogonal polynomials},
6136: J.\ Concr.\ Appl.\ Math. \textbf{3} (2005), 313--334.
6137: %
6138: \bibitem{CMV07} M.~J.~Cantero, L.~Moral, and L.~Vel\'azquez,
6139: \emph{Matrix orthogonal polynomials whose derivatives are also orthogonal},
6140: J.\ Approx.\ Theory \textbf{146} (2007), 174--211.
6141: %
6142: \bibitem{Cartan} E.~Cartan,
6143: \emph{Sur les domaines born\'es homog\`enes de l'espace de $n$
6144: variables complexes},
6145: Abh.\ Math.\ Sem.\ Univ.\ Hamburg \textbf{11} (1935), 116--162.
6146: %
6147: \bibitem{CG05} M.~Castro and F.~A.~Gr\"unbaum,
6148: \emph{Orthogonal matrix polynomials satisfying first order
6149: differential equations: a collection of instructive examples},
6150: J.\ Nonlinear Math.\ Physics \textbf{12} (2005), 63--76.
6151: %
6152: \bibitem{CG} M.~Castro and F.~A.~Gr\"unbaum,
6153: \emph{The algebra of differential operators associated to a family
6154: of matrix valued orthogonal polynomials: five instructive examples},
6155: Int.\ Math.\ Res.\ Not. \textbf{2006:7} (2006), 1--33.
6156: %
6157: \bibitem{CG07} M.~Castro and F.~A.~Gr\"unbaum,
6158: \emph{The noncommutative bispectral problem for operators of order one},
6159: Constr.\ Approx. \textbf{27} (2008), 329--347.
6160: %
6161: \bibitem{CH} G.~Chen and Y.~Hu,
6162: \emph{The truncated Hamburger matrix moment problems in the
6163: nondegenerate and degenerate cases, and matrix continued fractions},
6164: Linear Algebra Appl. \textbf{277} (1998), 199--236.
6165: %
6166: \bibitem{Chi} T.~S.~Chihara,
6167: \textit{An Introduction to Orthogonal Polynomials},
6168: Mathematics and Its Applications \textbf{13}, Gordon and Breach,
6169: New York-London-Paris (1978).
6170: %
6171: \bibitem{CG01} S.~Clark and F.~Gesztesy,
6172: \emph{Weyl--Titchmarsh $M$-function asymptotics for matrix-valued Schr\"odinger operators},
6173: Proc.\ London Math.\ Soc. \textbf{82} (2001), 701--724.
6174: %
6175: \bibitem{CG02} S.~Clark and F.~Gesztesy,
6176: \emph{Weyl--Titchmarsh $M$-function asymptotics, local uniqueness results, trace formulas,
6177: and Borg-type theorems for Dirac operators},
6178: Trans.\ Amer.\ Math.\ Soc. \textbf{354} (2002), 3475--3534.
6179: %
6180: \bibitem{CG03} S.~Clark and F.~Gesztesy,
6181: \emph{On Povzner--Wienholtz-type self-adjointness results for matrix-valued
6182: Sturm--Liouville operators},
6183: Proc.\ Math.\ Soc.\ Edinburgh \textbf{133A} (2003), 747--758.
6184: %
6185: \bibitem{CG04} S.~Clark and F.~Gesztesy,
6186: \emph{On Weyl--Titchmarsh theory for singular finite difference Hamiltonian systems},
6187: J.\ Comput.\ Appl.\ Math.\ {\bf 171} (2004), 151--184.
6188: %
6189: \bibitem{CGHL00} S.~Clark, F.~Gesztesy, H.~Holden, and B.~M.~Levitan,
6190: \emph{Borg-type theorems for matrix-valued Schr\"odinger operators},
6191: J.\ Diff.\ Eqs. \textbf{167} (2000), 181--210.
6192: %
6193: \bibitem{CGR} S.~Clark, F.~Gesztesy, and W.~Renger,
6194: \emph{Trace formulas and Borg-type theorems for matrix-valued Jacobi
6195: and Dirac finite difference operators},
6196: J.\ Diff.\ Eqs. \textbf{219} (2005), 144--182.
6197: 
6198: \bibitem{CGZ07} S.~Clark, F.~Gesztesy, and M.~Zinchenko,
6199: \emph{Weyl--Titchmarsh theory and Borg--Marchenko-type uniqueness results
6200: for CMV operators with matrix-valued Verblunsky coefficients},
6201: Operators and Matrices \textbf{1} (2007), 535--592.
6202: %
6203: \bibitem{CrS} W.~Craig and B.~Simon,
6204: \emph{Log H\"older continuity of the integrated density of states for
6205: stochastic Jacobi matrices},
6206: Comm.\ Math.\ Phys. \textbf{90} (1983), 207--218.
6207: %
6208: \bibitem{DKacta} D.~Damanik and R.~Killip,
6209: \emph{Half-line Schr\"odinger operators with no bound states},
6210: Acta Math. \textbf{193} (2004), 31--72.
6211: %
6212: \bibitem{DKSppt} D.~Damanik, R.~Killip, and B.~Simon,
6213: \emph{Perturbations of orthogonal polynomials with periodic recursion
6214: coefficients}, preprint.
6215: %
6216: \bibitem{Dean-Martin} P.~Dean and J.~L.~Martin,
6217: \emph{A method for determining the frequency spectra of disordered lattices in two-dimensions},
6218: Proc.\ Roy.\ Soc. (London) A \textbf{259} (1960), 409--418.
6219: %
6220: \bibitem{DGIM} A.~M.~Delgado, J.~S.~Geronimo, P.~Iliev, and F.~Marcell\'an,
6221: \emph{Two variable orthogonal polynomials and structured matrices},
6222: SIAM J.\ Matrix Anal.\ Appl. \textbf{28} (2006), 118--147 [electronic].
6223: %
6224: \bibitem{DG92} P.~Delsarte and Y.~V.~Genin,
6225: \emph{On a generalization of the Szeg\H{o}--Levinson recurrence and its
6226: application in lossless inverse scattering},
6227: IEEE Trans.\ Inform.\ Theory {\bf 38} (1992), 104--110.
6228: %
6229: \bibitem{DGK} P.~Delsarte, Y.~V.~Genin, and Y.~G.~Kamp,
6230: \emph{Orthogonal polynomial matrices on the unit circle},
6231: IEEE Trans.\ Circuits and Systems {\bf CAS-25} (1978), 149--160.
6232: %
6233: \bibitem{DGK3} P.~Delsarte, Y.~V.~Genin, and Y.~G.~Kamp,
6234: \emph{Planar least squares inverse polynomials, I.\ Algebraic properties},
6235: IEEE Trans.\ Circuits and Systems {\bf CAS-26} (1979), 59--66.
6236: %
6237: \bibitem{DGK79} P.~Delsarte, Y.~V.~Genin, and Y.~G.~Kamp,
6238: \emph{The Nevanlinna--Pick problem for matrix-valued functions},
6239: SIAM J.\ Appl.\ Math. {\bf 36} (1979), 47--61.
6240: %
6241: \bibitem{DGK2} P.~Delsarte, Y.~V.~Genin, and Y.~G.~Kamp,
6242: \emph{Schur parametrization of positive definite block-Toeplitz systems},
6243: SIAM J.\ Appl.\ Math. {\bf 36} (1979), 34--46.
6244: %
6245: \bibitem{DGK81} P.~Delsarte, Y.~V.~Genin, and Y.~G.~Kamp,
6246: \emph{Generalized Schur representation of matrix-valued functions},
6247: SIAM J.\ Algebraic Discrete Methods {\bf 2} (1981), 94--107.
6248: %
6249: \bibitem{Den} S.~Denisov,
6250: \emph{On Rakhmanov's theorem for Jacobi matrices},
6251: Proc.\ Amer.\ Math.\ Soc. \textbf{132} (2004), 847--852.
6252: %
6253: \bibitem{DS} H.~Dette and W.~J.~Studden,
6254: \emph{Matrix measures, moment spaces and Favard's theorem for the
6255: interval $[0,1]$ and $[0,\infty)$},
6256: Linear Algebra Appl. \textbf{345} (2002), 169--193.
6257: %
6258: \bibitem{DFK} V.~Dubovoj, B.~Fritzsche, and B.~Kirstein,
6259: \textit{Matricial Version of the Classical Schur Problem},
6260: Teubner Texts in Mathematics \textbf{129}, Teubner Verlag, Stuttgart (1992).
6261: %
6262: \bibitem{Dur1} A.~J.~Dur\'an,
6263: \emph{A generalization of Favard's theorem for polynomials satisfying a
6264: recurrence relation},
6265: J.\ Approx.\ Theory \textbf{74} (1993), 83--109.
6266: %
6267: \bibitem{Dur95} A.~J.~Dur\'an,
6268: \emph{On orthogonal polynomials with respect to a positive definite
6269: matrix of measures},
6270: Canad.\ J.\ Math. \textbf{47} (1995),  88--112.
6271: %
6272: \bibitem{Dur96} A.~J.~Dur\'an,
6273: \emph{Markov's theorem for orthogonal matrix polynomials},
6274: Canad.\ J.\ Math. \textbf{48} (1996),  1180--1195.
6275: %
6276: \bibitem{Dur6} A. J. Dur\'an,
6277: \emph{Matrix inner product having a matrix symmetric second order
6278: differential operator},
6279: Rocky Mountain J.\ Math. \textbf{27} (1997), 585--600.
6280: %
6281: \bibitem{Dur99} A.~J.~Dur\'an,
6282: \emph{Ratio asymptotics for orthogonal matrix polynomials},
6283: J.\ Approx.\ Theory \textbf{100} (1999), 304--344.
6284: %
6285: \bibitem{Dur7} A.~J.~Dur\'an,
6286: \emph{How to find weight matrices having symmetric
6287: second order differential operators with matrix leading
6288: coefficient}, preprint.
6289: %
6290: \bibitem{Dur8} A.~J.~Dur\'an,
6291: \emph{Generating weight matrices having symmetric
6292: second order differential operators from a trio of upper
6293: triangular matrices}, preprint.
6294: %
6295: \bibitem{DuDa} A.~J.~Dur\'an and E.~Daneri,
6296: \emph{Ratio asymptotics for orthogonal matrix polynomials with
6297: unbounded recurrence coefficients},
6298: J.\ Approx.\ Theory \textbf{110} (2001), 1--17.
6299: %
6300: \bibitem{DD} A.~J.~Dur\'an and E.~Daneri,
6301: \emph{Weak convergence for orthogonal matrix polynomials},
6302: Indag.\ Math.\ (N.S.) \textbf{13} (2002), 47--62.
6303: %
6304: \bibitem{DDe} A.~J.~Dura\'n and E.~Defez,
6305: \emph{Orthogonal matrix polynomials and quadrature formulas},
6306: Linear Algebra Appl. \textbf{345} (2002), 71--84.
6307: %
6308: \bibitem{DG04} A.~J.~Dur\'an and F.~A.~Gr\"unbaum,
6309: \emph{Orthogonal matrix polynomials satisfying second order
6310: differential equations},
6311: Int.\ Math.\ Res.\ Not. \textbf{2004:10} (2004), 461--484.
6312: %
6313: \bibitem{DG05a} A.~J.~Dur\'an and F.~A.~Gr\"unbaum,
6314: \emph{A survey on orthogonal matrix polynomials satisfying second
6315: order differential equations},
6316: J.\ Comput.\ Appl.\ Math. \textbf{178} (2005), 169--190.
6317: %
6318: \bibitem{DG05b} A.~J.~Dur\'an and F.~A.~Gr\"unbaum,
6319: \emph{Structural formulas for orthogonal matrix polynomials
6320: satysfying second order differential equations, I},
6321: Constr.\ Approx. \textbf{22} (2005), 255--271.
6322: %
6323: \bibitem{DG05c} A.~J.~Dur\'an and F.~A.~Gr\"unbaum,
6324: \emph{Orthogonal matrix polynomials, scalar type Rodrigues'
6325: formulas and Pearson equations}, J.\ Approx.\ Theory. \textbf{134}
6326: (2005), 267--280.
6327: %
6328: \bibitem{DG05d} A.~J.~Dur\'an and F.~A.~Gr\"unbaum,
6329: \emph{A characterization for a class of weight matrices with orthogonal
6330: matrix polynomials satisfying second order differential equations},
6331: Int.\ Math.\ Res.\ Not.\ \textbf{23} (2005), 1371--1390.
6332: %
6333: \bibitem{DG06} A.~J.~Dur\'an and F.~A.~Gr\"unbaum,
6334: \emph{P.~A.~M.~Dirac meets M.~G.~Krein: matrix orthogonal
6335: polynomials and Dirac´s equation},
6336: J.\ Phys.\ A: Math.\ Gen. \textbf{39} (2006), 3655--3662.
6337: %
6338: \bibitem{DG07a} A.~J.~Dur\'an and F.~A.~Gr\"unbaum,
6339: \emph{Matrix orthogonal polynomials satisfying second-order
6340: differential equations: Coping without help from group
6341: representation theory},
6342: J.\ Approx.\ Theory \textbf{148} (2007), 35--48.
6343: %
6344: \bibitem{DG07b} A.~J.~Dur\'an and F.~A.~Gr\"unbaum,
6345: \emph{Matrix differential equations and scalar polynomials
6346: satisfying higher order recursions}, preprint.
6347: %
6348: \bibitem{DGPT}
6349: A.~ Duran, F.~A.~Gr\"unbaum, I.~Pacharoni, and J.~A.~Tirao,
6350: \emph{Matrix orthogonal polynomials and differential operators,
6351: with examples and applications}, in preparation.
6352: %
6353: \bibitem{DdI07a} A.~J.~Dur\'an and M.~D.~de la Iglesia,
6354: \emph{Some examples of orthogonal matrix polynomials satisfying
6355: odd order differential equations},
6356: to appear in J. Approx. Theory.
6357: %
6358: \bibitem{DdI07b} A.~J.~Dur\'an and M.~D.~de la Iglesia,
6359: \emph{Second order differential operators having several families
6360: of orthogonal matrix polynomials as eigenfunctions}, preprint.
6361: %
6362: \bibitem{DL} A.~J.~Dur\'an and P.~L\'opez-Rodr\'\i guez,
6363: \emph{Orthogonal matrix polynomials: zeros and Blumenthal's theorem},
6364: J.\ Approx.\ Theory \textbf{84} (1996), 96--118.
6365: %
6366: \bibitem{DL97a} A.~J.~Dur\'an and P.~L\'opez-Rodr\'\i guez,
6367: \emph{The $L^p$ space of a positive definite matrix of measures
6368: and density of matrix polynomials in $L^1$},
6369: J.\ Approx.\ Theory \textbf{90} (1997), 299--318.
6370: %
6371: \bibitem{DL97b} A.~J.~Dur\'an and P.~L\'opez-Rodr\'\i guez,
6372: \emph{Density questions for the truncated matrix moment problem},
6373: Canad.\ J.\ Math.\ \textbf{49} (1997), 708--721.
6374: %
6375: \bibitem{DL00} A.~J.~Dur\'an and P.~L\'opez-Rodr\'\i guez,
6376: \emph{$N$-extremal matrices of measures for an indeterminate
6377: matrix moment problem},
6378: J.\ Funct.\ Anal.\ \textbf{174} (2000), 301--321.
6379: %
6380: \bibitem{DuLo4} A.~J.~Dur\'an and P.~L\'opez-Rodr\'\i guez,
6381: \emph{The matrix moment problem},
6382: Margarita Mathematica en memoria de Jos\'e Javier Guadalupe
6383: (L.~Espa\~nol and J.~L.~Varona, eds.), pp.~333--348,
6384: Universidad de La Rioja, Logro\~no (2001).
6385: %
6386: \bibitem{DL04} A.~J.~Dur\'an and P.~L\'opez-Rodr\'\i guez,
6387: \emph{Orthogonal matrix polynomials}, Laredo's SIAG Lecture Notes,
6388: (R.~\'Alvarez-Nodarse et al., eds.), pp.~11--43, Advances in the Theory of
6389: Special Functions and Orthogonal Polynomials, Nova Science
6390: Publishers \textbf{1} (2003).
6391: %
6392: \bibitem{DuLo5} A.~J.~Dur\'an and P. L\'opez-Rodr\'\i guez,
6393: \emph{Structural formulas for orthogonal matrix polynomials satisfying second order
6394: differential equations, I},
6395: Constr.\ Approx. \textbf{26} (2007), 29--47.
6396: %
6397: \bibitem{DL07} A.~J.~Dur\'an and P. L\'opez-Rodr\'\i guez,
6398: \emph{Structural formulas for orthogonal matrix polynomials
6399: satisfying second order differential equations, II},
6400: Constr.\ Approx.\ \textbf{26} (2007) 29--47.
6401: %
6402: \bibitem{DLS} A.~J.~Dur\'an, P.~L\'opez-Rodr\'\i guez, and E.~B.~Saff,
6403: \emph{Zero asymptotic behaviour for orthogonal matrix polynomials},
6404: J.\ d'Analyse Math. \textbf{78} (1999), 37--60.
6405: %
6406: \bibitem{DP02} A.~J.~Dur\'an and B.~Polo,
6407: \emph{Gaussian quadrature formulae for matrix weights},
6408: Linear Algebra Appl. \textbf{355} (2002), 119--146.
6409: %
6410: \bibitem{DP} A.~J.~Dur\'an and B.~Polo,
6411: \emph{Matrix Christoffel functions},
6412: Constr.\ Approx. \textbf{20} (2004), 353--376.
6413: %
6414: \bibitem{DVA} A.~J.~Dur\'an and W.~Van Assche,
6415: \emph{Orthogonal matrix polynomials and higher-order recurrence relations},
6416: Linear Algebra Appl. \textbf{219} (1995), 261--280.
6417: %
6418: \bibitem{Dym} H.~Dym,
6419: \textit{$J$ Contractive Matrix Functions, Reproducing Kernel Hilbert
6420: Spaces and Interpolation},
6421: CBMS Regional Conference Series in Math. \textbf{71},
6422: American Mathematical Society, Providence, R.I. (1989).
6423: %
6424: \bibitem{Dyu} Yu.~M.~Dyukarev,
6425: \emph{Deficiency numbers of symmetric operators generated by block Jacobi matrices},
6426: Sb.\ Math. \textbf{197} (2006), 1177--1203.
6427: %
6428: \bibitem{ENZG91} T.~Erd\'elyi, P.~Nevai, J.~Zhang, and J.~Geronimo,
6429: \emph{A simple proof of ``Favard's theorem" on the unit circle},
6430: Atti Sem.\ Mat.\ Fis.\ Univ.\ Modena {\bf 39} (1991), 551--556.
6431: %
6432: \bibitem{FG99} L.~Faybusovich and M.~Gekhtman,
6433: \emph{On Schur flows},
6434: J.\ Phys.\ A {\bf 32} (1999), 4671--4680.
6435: %
6436: \bibitem{FrB} G.~Freud,
6437: \textit{Orthogonal Polynomials},
6438: Pergamon Press, Oxford-New York (1971).
6439: %
6440: \bibitem{Fu} P.~A.~Fuhrmann,
6441: \emph{Orthogonal matrix polynomials and system theory},
6442: Rend.\ Sem.\ Mat.\ Univ.\ Politec.\ Torino \textbf{1987}, Special
6443: Issue, 68--124.
6444: %
6445: \bibitem{Ger81} J.~S.~Geronimo,
6446: \emph{Matrix orthogonal polynomials on the unit circle},
6447: J.\ Math.\ Phys. \textbf{22} (1981), 1359--1365.
6448: %
6449: \bibitem{Ger82} J.~S.~Geronimo,
6450: \emph{Scattering theory and matrix orthogonal polynomials on the real line},
6451: Circuits Systems Signal Process \textbf{1} (1982), 471--495.
6452: %
6453: \bibitem{Ge92} J.~S.~Geronimo,
6454: \emph{Polynomials orthogonal on the unit circle with random recurrence coefficients},
6455: Methods of Approximation Theory in Complex Analysis and Mathematical Physics
6456: (Leningrad, 1991), pp.~43--61,
6457: Lecture Notes in Math. \textbf{1550}, Springer, Berlin (1993).
6458: %
6459: \bibitem{GC} J.~S.~Geronimo and K.~M.~Case,
6460: \emph{Scattering theory and polynomials orthogonal on the unit circle},
6461: J.\ Math.\ Phys. {\bf 20} (1979), 299--310.
6462: %
6463: \bibitem{GW07} J.~S.~Geronimo and H.~Woerdeman,
6464: \emph{Two variable orthogonal polynomials on the bicircle and
6465: structured matrices},
6466: SIAM J.\ Matrix Anal.\ Appl. \textbf{29} (2007), 796--825.
6467: %
6468: \bibitem{Ger44} Ya.~L.~Geronimus,
6469: \emph{On polynomials orthogonal on the circle, on trigonometric
6470: moment problem, and on allied Carath\'eodory and Schur functions},
6471: Mat.\ Sb. \textbf{15} (1944), 99--130 [Russian].
6472: %
6473: \bibitem{Ger46} Ya.~L.~Geronimus,
6474: \emph{On the trigonometric moment problem},
6475: Ann.\ of Math. (2) {\bf 47} (1946), 742--761.
6476: %
6477: \bibitem{GBk1} Ya.~L.~Geronimus,
6478: \textit{Polynomials Orthogonal on a Circle and Their Applications},
6479: Amer.\ Math.\ Soc.\ Translation \textbf{1954} (1954), no.~104, 79 pp.
6480: %
6481: \bibitem{GBk} Ya.~L.~Geronimus,
6482: \textit{Orthogonal Polynomials: Estimates, Asymptotic Formulas,
6483: and Series of Polynomials Orthogonal on the Unit Circle and on an
6484: Interval},
6485: Consultants Bureau, New York (1961).
6486: %
6487: \bibitem{Ger77} Ya.~L.~Geronimus,
6488: \emph{Orthogonal polynomials},
6489: Engl.\ translation of the appendix to the Russian
6490: translation of Szeg\H{o}'s book \cite{Szb}, in ``Two Papers on
6491: Special Functions," Amer.\ Math.\ Soc.\ Transl., Ser.~2, Vol.~108,
6492: pp.~37--130, American Mathematical Society, Providence, R.I. (1977).
6493: %
6494: \bibitem{GKMT} F.~Gesztesy, N.~J.~Kalton, K.~A.~Makarov, and E.~Tsekanovskii,
6495: \emph{Some applications of operator-valued Herglotz functions},
6496: Operator Theory, System Theory and Related Topics (Beer-Sheva/Rehovot, 1997),
6497: pp.~271--321, Oper.\ Theory Adv.\ Appl. \textbf{123}, Birkh\"auser, Basel (2001).
6498: %
6499: \bibitem{GKM} F.~Gesztesy, A.~Kiselev, and K.~Makarov,
6500: \emph{Uniqueness results for matrix-valued Schr\"odinger, Jacobi, and
6501: Dirac-type operators},
6502: Math.\ Nachr. \textbf{239/240} (2002), 103--145.
6503: %
6504: \bibitem{GS03} F.~Gesztesy and L.~A.~Sakhnovich,
6505: \emph{A class of matrix-valued Schr\"odinger operators with
6506: prescribed finite-band spectra},
6507: Reproducing Kernel Hilbert Spaces, Positivity, System Theory and
6508: Related Topics (D.~Alpay, ed.), pp.~213--253, Operator Theory: Advances and
6509: Applications \textbf{143}, Birkh\"auser, Basel (2003).
6510: %
6511: \bibitem{GS00} F.~Gesztesy and B.~Simon,
6512: \emph{On local Borg--Marchenko uniqueness results},
6513: Comm.\ Math.\ Phys. \textbf{211} (2000), 273--287.
6514: %
6515: \bibitem{GT} F.~Gesztesy and E.~Tsekanovskii,
6516: \emph{On matrix-valued Herglotz functions},
6517: Math.\ Nachr. \textbf{218} (2000), 61--138.
6518: %
6519: \bibitem{GK} I.~Gohberg and M.~Krein,
6520: \textit{Introduction to the Theory of Linear Nonselfadjoint Operators},
6521: Translations of Mathematical Monographs \textbf{18},
6522: American Mathematical Society, Providence, R.I. (1969).
6523: %
6524: \bibitem{Gol02} L.~Golinskii,
6525: \emph{Quadrature formula and zeros of para-orthogonal polynomials on
6526: the unit circle},
6527: Acta Math.\ Hungar.\ {\bf 96} (2002), 169--186.
6528: %
6529: \bibitem{KhGo} L.~Golinskii and S.~Khrushchev,
6530: \emph{Ces\`aro asymptotics for orthogonal polynomials on the unit
6531: circle and classes of measures},
6532: J.\ Approx.\ Theory \textbf{115} (2002), 187--237.
6533: %
6534: \bibitem{Gru} F.~A.~Gr\"unbaum,
6535: \emph{Matrix valued Jacobi polynomials},
6536: Bull.\ Sciences Math. \textbf{127} (2003), 207--214.
6537: %
6538: \bibitem{Gru07} F.~A.~Gr\"unbaum,
6539: \emph{Random walks and orthogonal polynomials: some challenges},
6540: preprint.
6541: %
6542: \bibitem{GI07a} F.~A.~Gr\"unbaum and M.~D.~de la Iglesia,
6543: \emph{Matrix valued orthogonal polynomials related to SU$(N+1)$,
6544: their algebras of differential operators and the corresponding curves},
6545: Exp.\ Math.\ \textbf{16} (2007), 189--207.
6546: %
6547: \bibitem{GI07b} F.~A.~Gr\"unbaum and M.~D.~de la Iglesia,
6548: \emph{Matrix valued orthogonal polynomials arising from group
6549: representation theory and a family of quasi-birth-and-death processes},
6550: preprint.
6551: %
6552: \bibitem{GI03} F.~A.~Gr\"unbaum and P.~Iliev,
6553: \emph{A noncommutative version of the bispectral problem},
6554: J.\ Comput.\ Appl.\ Math.\ \textbf{161} (2003), 99--118.
6555: %
6556: \bibitem{GPT01} F.~A.~Gr\"unbaum, I.~Pacharoni, and J.~A.~Tirao,
6557: \emph{A matrix valued solution to Bochner's problem},
6558: J.\ Phys.\ A: Math.\ Gen. \textbf{34} (2001), 10647--10656.
6559: %
6560: \bibitem{GPT02a} F.~A.~Gr\"unbaum, I.~Pacharoni, and J.~A.~Tirao,
6561: \emph{Matrix valued spherical functions associated to the three
6562: dimensional hyperbolic space},
6563: Internat.\ J.\ of Math. \textbf{13} (2002), 727--784.
6564: %
6565: \bibitem{GPT02} F.~A.~Gr\"unbaum, I.~Pacharoni, and J.~A.~Tirao,
6566: \emph{Matrix valued spherical functions associated to the complex projective plane},
6567: J.\ Funct.\ Anal. {\bf 188} (2002), 350--441.
6568: %
6569: \bibitem{GPT03} F.~A.~Gr\"unbaum, I.~Pacharoni, and J.~A.~Tirao,
6570: \emph{Matrix valued orthogonal polynomials of the Jacobi type},
6571: Indag.\ Math. \textbf{14} (2003), 353--366.
6572: %
6573: \bibitem{GPT04} F.~A.~Gr\"unbaum, I.~Pacharoni, and J.~A.~Tirao,
6574: \emph{An invitation to matrix-valued spherical functions: Linearization of
6575: products in the case of complex projective space $P\sb 2({\mathbb C})$},
6576: Modern Signal Processing, pp.~147--160,
6577: Math.\ Sci.\ Res.\ Inst.\ Publ. \textbf{46},
6578: Cambridge University Press, Cambridge (2004).
6579: %
6580: \bibitem{GPT05} F.~A.~Gr\"unbaum, I.~Pacharoni, and J.~A.~Tirao,
6581: \emph{Matrix valued orthogonal polynomials of the Jacobi type:
6582: The role of group representation theory},
6583: Ann.\ Inst.\ Fourier (Grenoble) \textbf{55} (2005), 2051--2068.
6584: %
6585: \bibitem{GT07} F.~A.~Gr\"unbaum and J.~A.~Tirao,
6586: \emph{The algebra of differential operators associated to a weight matrix},
6587: Integr.\ Equ.\ Oper.\ Th. \textbf{58} (2007), 449--475.
6588: %
6589: \bibitem{Helg} S.~Helgason,
6590: \textit{Differential Geometry, Lie Groups, and Symmetric Spaces},
6591: corrected reprint of the 1978 original, Graduate Studies in
6592: Mathematics \textbf{34},
6593: American Mathematical Society, Providence, R.I. (2001).
6594: %
6595: \bibitem{Helms} L.~L.~Helms,
6596: \textit{Introduction to Potential Theory},
6597: Pure and Applied Mathematics {\bf 22},
6598: Wiley-Interscience, New York (1969).
6599: %
6600: \bibitem{HL1} H.~Helson and D.~Lowdenslager,
6601: \emph{Prediction theory and Fourier series in several variables},
6602: Acta Math. \textbf{99} (1958), 165--202.
6603: %
6604: \bibitem{HL2} H.~Helson and D.~Lowdenslager,
6605: \emph{Prediction theory and Fourier series in several variables, II},
6606: Acta Math. {\bf 106} (1961), 175--213.
6607: %
6608: \bibitem{IsmBk} M.~E.~H.~Ismail,
6609: \textit{Classical and Quantum Orthogonal Polynomials in One Variable},
6610: Encyclopedia of Mathematics and its Applications \textbf{98},
6611: Cambridge University Press, Cambridge (2005).
6612: %
6613: \bibitem{JNT89} W.~B.~Jones, O.~Nj\r{a}stad, and W.~J.~Thron,
6614: \emph{Moment theory, orthogonal polynomials, quadrature, and
6615: continued fractions associated with the unit circle},
6616: Bull.\ London Math.\ Soc.\ \textbf{21} (1989), 113--152.
6617: %
6618: \bibitem{Kai74} T.~Kailath,
6619: \emph{A view of three decades of linear filtering theory},
6620: IEEE Trans.\ Inform.\ Theory \textbf{IT-20} (1974), 146--181.
6621: %
6622: \bibitem{Kai87} T.~Kailath,
6623: \emph{Signal processing applications of some moment problems},
6624: Moments in Mathematics, (San Antonio, Tex., 1987),
6625: Proc.\ Sympos.\ Appl.\ Math. \textbf{37}, pp.~71--109,
6626: American Mathematical Society, Providence, R.I. (1987).
6627: %
6628: \bibitem{K6.1} T.~Kailath,
6629: \emph{Norbert Wiener and the development of mathematical engineering},
6630: The Legacy of Norbert Wiener: A Centennial Symposium, Proc.\ Sympos.\
6631: Pure Math. \textbf{60}, pp.~93--116, American Mathematical Society,
6632: Providence, R.I. (1997).
6633: %
6634: \bibitem{KVM} T.~Kailath, A.~Vieira, and M.~Morf,
6635: \emph{Inverses of Toeplitz operators, innovations, and orthogonal
6636: polynomials},
6637: SIAM Rev. \textbf{20} (1978), 106--119.
6638: %
6639: \bibitem{Kato} T.~Kato,
6640: \textit{Perturbation Theory for Linear Operators},
6641: second edition, Grundlehren der Mathematischen Wissenschaften
6642: \textbf {132}, Springer, Berlin-New York (1976).
6643: %
6644: \bibitem{Kh2000} S.~Khrushchev,
6645: \emph{Schur's algorithm, orthogonal polynomials, and convergence of
6646: Wall's continued fractions in $L^2 (\bbT)$},
6647: J.\ Approx.\ Theory \textbf{108} (2001), 161--248.
6648: %
6649: \bibitem{Khr} S.~Khrushchev,
6650: \emph{Classification theorems for general orthogonal polynomials on
6651: the unit circle},
6652: J.\ Approx.\ Theory \textbf{116} (2002), 268--342.
6653: %
6654: \bibitem{KN} R.\ Killip and I.\ Nenciu,
6655: \emph{Matrix models for circular ensembles},
6656: Int.\ Math.\ Res.\ Not. \textbf{2004:50} (2004), 2665--2701.
6657: %
6658: \bibitem{KS} R.~Killip and B.~Simon,
6659: \emph{Sum rules for Jacobi matrices and their applications to spectral theory},
6660: Ann.\ of Math.\ {\bf 158} (2003), 253--321.
6661: %
6662: \bibitem{Kol} A.~N.~Kolmogorov,
6663: \emph{Stationary sequences in Hilbert space},
6664: Bull.\ Univ.\ Moscow {\bf 2} (1941), 40 pp. [Russian].
6665: %
6666: \bibitem{KS88} S.~Kotani and B.~Simon,
6667: \emph{Stochastic Schr\"odinger operators and Jacobi matrices on the strip},
6668: Comm.\ Math.\ Phys. \textbf{119} (1988), 403--429.
6669: %
6670: \bibitem{Krein1} M.~G.~Krein,
6671: \emph{On a generalization of some investigations of G.~Szeg\H{o},
6672: W.~M.~Smirnov, and A.~N.~Kolmogorov},
6673: Dokl.\ Akad.\ Nauk SSSR \textbf{46} (1945), 91--94.
6674: %
6675: \bibitem{Krein2} M.~G.~Krein,
6676: \emph{On a problem of extrapolation of A.~N.~Kolmogorov},
6677: Dokl.\ Akad.\ Nauk SSSR {\bf 46} (1945), 306--309.
6678: %
6679: \bibitem{Krein3} M.~G.~Krein,
6680: \emph{Infinite $J$-matrices and a matrix moment problem},
6681: Dokl.\ Akad.\ Nauk SSSR \textbf{69} (1949), 125--128 [Russian].
6682: %
6683: \bibitem{Lac} J.~Lacroix,
6684: \emph{The random Schr\"odinger operator in a strip},
6685: Probability Measures on Groups, VII (Oberwolfach, 1983), pp.~280--297,
6686: Lecture Notes in Math. \textbf{1064}, Springer, Berlin (1984).
6687: %
6688: \bibitem{Landau} H.~J.~Landau,
6689: \emph{Maximum entropy and the moment problem},
6690: Bull.\ Amer.\ Math.\ Soc. \textbf{16} (1987), 47--77.
6691: %
6692: \bibitem{Land} N.~S.~Landkof,
6693: \textit{Foundations of Modern Potential Theory},
6694: Springer-Verlag, Berlin-New York (1972).
6695: %
6696: \bibitem{LR} L.~Lerer and A.~C.~M.~Ran,
6697: \emph{A new inertia theorem for Stein equations, inertia of invertible
6698: Hermitian block Toeplitz matrices and matrix orthogonal polynomials},
6699: Integr.\ Equ.\ Oper.\ Th. \textbf{47} (2003), 339--360.
6700: %
6701: \bibitem{Lev} N.~Levinson,
6702: \emph{The Wiener RMS (root-mean square) error criterion in filter
6703: design and prediction},
6704: J.\ Math.\ Phys.\ Mass.\ Inst.\ Tech. \textbf{25} (1947), 261--278.
6705: %
6706: \bibitem{Lop} P.~L\'opez-Rodr\'\i guez,
6707: \emph{Riesz's theorem for orthogonal matrix polynomials},
6708: Constr.\ Approx. \textbf{15} (1999),  135--151.
6709: %
6710: \bibitem{Lop01} P.~L\'opez-Rodr\'\i guez,
6711: \emph{Nevanlinna parametrization for a matrix moment problem},
6712: Math.\ Scand. \textbf{89} (2001), 245--267.
6713: %
6714: \bibitem{Lub} D.~S.~Lubinsky,
6715: \emph{A new approach to universality limits involving orthogonal polynomials},
6716: to appear in Ann.\ of Math.
6717: %
6718: \bibitem{MarRod} F.~Marcell\'an and I.~Rodriguez,
6719: \emph{A class of matrix orthogonal polynomials on the unit circle},
6720: Linear Algebra Appl. \textbf{121}, (1989), 233--241.
6721: %
6722: \bibitem{MS} F.~Marcell\'an and G.~Sansigre,
6723: \emph{On a class of matrix orthogonal polynomials on the real line},
6724: Linear Algebra Appl. \textbf{181} (1993), 97--109.
6725: %
6726: \bibitem{MY} F.~Marcell\'an and H.~O.~Yakhlef,
6727: \emph{Recent trends on analytic properties of matrix orthonormal polynomials},
6728: Orthogonal Polynomials, Approximation Theory, and Harmonic Analysis
6729: (Inzel, 2000), Electron.\ Trans.\ Numer.\ Anal. \textbf{14} (2002), 127--141
6730: [electronic].
6731: %
6732: \bibitem{MZ} F.~Marcell\'an and S.~M.~Zagorodnyuk,
6733: \emph{On the basic set of solutions of a high-order linear difference equation},
6734: J.\ Diff.\ Equ.\ Appl. \textbf{12} (2006), 213--228.
6735: %
6736: \bibitem{MN} A.~M\'at\'e and P.~Nevai,
6737: \emph{Bernstein's inequality in $L\sp{p}$ for $0<p<1$ and $(C,\,1)$
6738: bounds for orthogonal polynomials},
6739: Ann.\ of Math. (2) \textbf{111} (1980),  145--154.
6740: %
6741: \bibitem{MNT88} A.~M\'at\'e, P.~Nevai, and V.~Totik,
6742: \emph{Strong and weak convergence of orthogonal polynomials},
6743: Am.\ J.\ Math. \textbf{109} (1987), 239--281.
6744: %
6745: \bibitem{M05a} L.~Miranian,
6746: \emph{Matrix valued orthogonal polynomials},
6747: Ph.D.\ Thesis, University of California at Berkeley (2005).
6748: %
6749: \bibitem{M05b} L.~Miranian,
6750: \emph{Matrix valued orthogonal polynomials on the real line:
6751: Some extensions of the classical theory},
6752: J.\ Phys.\ A: Math.\ Gen. \textbf{38} (2005), 5731--5749.
6753: %
6754: \bibitem{M05c} L.~Miranian,
6755: \emph{On classical orthogonal polynomials and differential operators},
6756: J.\ Phys.\ A: Math.\ Gen. \textbf{38} (2005), 6379--6383.
6757: %
6758: \bibitem{MV} S.~Molchanov and B.~Vainberg,
6759: \emph{Schr\"odinger operators with matrix potentials. Transition from
6760: the absolutely continuous to the singular spectrum},
6761: J.\ Funct.\ Anal. \textbf{215} (2004),  111--129.
6762: %
6763: \bibitem{NevFr} P.~Nevai,
6764: \emph{G\'eza Freud, orthogonal polynomials and Christoffel functions. A case study},
6765: J.\ Approx.\ Theory {\bf 48} (1986), 167~pp.
6766: %
6767: \bibitem{Nik} E.~M.~Nikishin,
6768: \emph{The discrete Sturm--Liouville operator and some problems of function theory},
6769: J.\ Soviet Math. \textbf{35} (1987), 2679--2744; Russian original in
6770: Trudy Sem.\ Petrovsk. \textbf{10} (1984), 3--77.
6771: %
6772: \bibitem{PR} I.\ Pacharoni and P.\ Rom\'an,
6773: \emph{A sequence of matrix valued orthogonal polynomials
6774: associated to spherical functions},
6775: Constr.\ Approx. \textbf{28} (2008), 127--147.
6776: %
6777: \bibitem{PT04} I.~Pacharoni and J.~A.~Tirao,
6778: \emph{Three term recursion relation for spherical functions
6779: associated to the complex projective plane},
6780: Math.\ Phys.\ Anal.\ Geom. \textbf{7} (2004), 193--221.
6781: %
6782: \bibitem{PT07a} I.~Pacharoni and J.~A.~Tirao,
6783: \emph{Matrix valued orthogonal polynomials arising from the
6784: complex projective space},
6785: Constr.\ Approx.\  \textbf{25} (2007), 177--192.
6786: %
6787: \bibitem{PT07b} I.~Pacharoni and J.~A.~Tirao,
6788: \emph{Matrix valued orthogonal polynomials associated to the group SU$(N)$},
6789: in preparation.
6790: %
6791: \bibitem{Ran} T.~Ransford,
6792: \textit{Potential Theory in the Complex Plane},
6793: Press Syndicate of the Univesity of Cambridge, New York (1995).
6794: %
6795: \bibitem{RS4} M.~Reed and B.~Simon,
6796: \textit{Methods of Modern Mathematical Physics, IV: Analysis of Operators},
6797: Academic Press, New York (1978).
6798: %
6799: \bibitem{Rod90} L.~Rodman,
6800: \emph{Orthogonal matrix polynomials},
6801: Orthogonal Polynomials (Columbus, OH, 1989), pp.~345--362,
6802: NATO Adv.\ Sci.\ Inst.\ Ser.\ C Math.\ Phys.\ Sci. \textbf{294},
6803: Kluwer, Dordrecht (1990).
6804: %
6805: \bibitem{RT} P.~Rom\'an and J.~A.~Tirao,
6806: \emph{Spherical functions, the complex hyperbolic plane and the
6807: hypergeometric operator},
6808: Internat.\ J.\ Math. \textbf{17} (2006), 1151--1173.
6809: %
6810: \bibitem{Ros} M.~Rosenberg,
6811: \emph{The square integrability of matrix valued functions with respect
6812: to a non-negative Hermitian measure},
6813: Duke Math.\ J. \textbf{31} (1964), 291--298.
6814: %
6815: \bibitem{Schur} I.~Schur,
6816: \emph{\"Uber Potenzreihen, die im Innern des Einheitskreises beschr
6817: \"ankt sind, I, II},
6818: J.\ Reine Angew.\ Math. \textbf{147} (1917), 205--232;
6819: \textbf{148} (1918), 122--145; English translation in
6820: ``I.~Schur Methods in Operator Theory and Signal Processing"
6821: (I.~Gohberg, ed.), pp.~31--59, 66--88, Operator Theory: Advances and Applications
6822: \textbf{18}, Birkh\"auser, Basel (1986).
6823: %
6824: \bibitem{SB04} H.~Schulz-Baldes,
6825: \emph{Perturbation theory for Lyapunov exponents of
6826: an Anderson model on a strip},
6827: Geom.\ Funct.\ Anal.\ \textbf{14} (2004), 1089--1117.
6828: %
6829: \bibitem{SB07} H.~Schulz-Baldes,
6830: \emph{Rotation numbers for Jacobi matrices with matrix entries},
6831: Math.\ Phys.\ Electron.\ J. \textbf{13} (2007), 40 pp.
6832: %
6833: \bibitem{ScZ} B.~Schwarz and A.~Zaks,
6834: \emph{Matrix M\"obius transformations},
6835: Comm.\ Algebra \textbf{9} (1981), 1913--1968.
6836: %
6837: \bibitem{S} B.~Simon,
6838: \emph{Orthogonal Polynomials on the Unit Circle. Part~1. Classical Theory},
6839: Colloquium Publications \textbf{54.1}, American Mathematical Society,
6840: Providence, R.I. (2005).
6841: %
6842: \bibitem{S2} B.~Simon,
6843: \emph{Orthogonal Polynomials on the Unit Circle. Part~2. Spectral Theory},
6844: Colloquium Publications \textbf{54.2}, American Mathematical Society,
6845: Providence, R.I. (2005).
6846: %
6847: \bibitem{S-TI} B.~Simon,
6848: \emph{Trace Ideals and Their Applications}, second edition,
6849: Mathematical Surveys and Monographs
6850: \textbf{120}, American Mathematical Society, Providence, R.I. (2005).
6851: %
6852: \bibitem{SimonRev} B.~Simon,
6853: \emph{CMV matrices: Five years after},
6854: J.\ Comput.\ Appl.\ Math. \textbf{208} (2007), 120--154.
6855: %
6856: \bibitem{EqMC} B.~Simon,
6857: \emph{Equilibrium measures and capacities in spectral theory},
6858: Inverse Problems and Imaging \textbf{1} (2007), 376-382.
6859: %
6860: \bibitem{S308} B.~Simon,
6861: \emph{Rank one perturbations and the zeros of paraorthogonal
6862: polynomials on the unit circle},
6863: J.\ Math.\ Anal.\ Appl. \textbf{329} (2007), 376--382.
6864: %
6865: \bibitem{2ext} B.~Simon,
6866: \emph{Two extensions of Lubinsky's universality theorem},
6867: to appear in J.\ Anal.\ Math.
6868: %
6869: \bibitem{Weak-cd} B.~Simon,
6870: \emph{Weak convergence of CD kernels and applications}, preprint.
6871: %
6872: \bibitem{Sim-cd} B.~Simon,
6873: \emph{The Christoffel--Darboux kernel},
6874: in preparation, to appear in Maz'ya Birthday volume.
6875: %
6876: \bibitem{Rice} B.~Simon,
6877: \textit{Szeg\H{o}'s Theorem and Its Descendants: Spectral Theory for
6878: $L^2$ Perturbations of Orthogonal Polynomials}, in preparation;
6879: to be published by Princeton University Press.
6880: %
6881: \bibitem{Sinap} A.~Sinap,
6882: \emph{Gaussian quadrature for matrix valued functions on the real line},
6883: J.\ Comput.\ Appl.\ Math. \textbf{65} (1995),  369--385.
6884: %
6885: \bibitem{SinapVanAsche} A.~Sinap and W.~Van Assche,
6886: \emph{Orthogonal matrix polynomials and applications},
6887: J.\ Comput.\ Appl.\ Math. \textbf{66} (1996), 27--52.
6888: %
6889: \bibitem{SVI} V.~N.~Sorokin and J.~Van Iseghem,
6890: \emph{Matrix continued fractions},
6891: J.\ Approx.\ Theory \textbf{96} (1999),  237--257.
6892: %
6893: \bibitem{StT} H.~Stahl and V.~Totik,
6894: \emph{General Orthogonal Polynomials},
6895: Encyclopedia of Mathematics and its Applications \textbf{43},
6896: Cambridge University Press, Cambridge (1992).
6897: %
6898: \bibitem{Stie} T.~Stieltjes,
6899: \emph{Recherches sur les fractions continues},
6900: Ann.\ Fac.\ Sci.\ Univ.\ Toulouse {\bf 8} (1894--1895), J76--J122;
6901: ibid. {\bf 9}, A5--A47.
6902: %
6903: \bibitem{Sz22a} G.~Szeg\H{o},
6904: \emph{\"Uber den asymptotischen Ausdruck von Polynomen, die durch eine
6905: Orthogonalit\"atseigenschaft definiert sind},
6906: Math.\ Ann. {\bf 86} (1922), 114--139.
6907: %
6908: \bibitem{Sz24} G.~Szeg\H{o},
6909: {\it Bemerkungen zu einer Arbeit von Herrn M. Fekete: \"Uber die
6910: Verteilung der Wurzeln bei gewissen algebraischen Gleichungen mit
6911: ganzzahligen Koeffizienten},
6912: Math.\ Z. {\bf 21} (1924), 203--208.
6913: %
6914: \bibitem{Szb} G. Szeg\H{o},
6915: \textit{Orthogonal Polynomials}, Amer.\ Math.\ Soc.\ Colloq.\ Publ.
6916: \textbf{23}, American Mathematical Society, Providence, R.I. (1939);
6917: third edition, (1967).
6918: %
6919: \bibitem{Teschl} G.~Teschl,
6920: \textit{Jacobi Operators and Completely Integrable Nonlinear Lattices},
6921: Mathematical Surveys and Monographs \textbf{72}, American
6922: Mathematical Society, Providence, R.I. (2000).
6923: %
6924: \bibitem{Tirao} J.~A.~Tirao,
6925: \emph{The matrix valued hypergeometric equation},
6926: Proc.\ Nat.\ Acad.\ Sci.\ U.S.A. \textbf{100} (2003), 8138--8141.
6927: %
6928: \bibitem{Tot} V.~Totik,
6929: \emph{Asymptotics for Christoffel functions for general measures on
6930: the real line},
6931: J.\ Anal.\ Math. \textbf{81} (2000), 283--303.
6932: %
6933: \bibitem{Tot-acta} V.~Totik,
6934: \emph{Polynomial inverse images and polynomial inequalities},
6935: Acta Math. \textbf{187} (2001), 139--160.
6936: %
6937: \bibitem{Tot-prep} V.~Totik,
6938: \emph{Universality and fine zero spacing on general sets},
6939: in preparation.
6940: %
6941: \bibitem{Ull} J.~L.~Ullman,
6942: \emph{On the regular behaviour of orthogonal polynomials},
6943: Proc.\ London Math.\ Soc. (3) \textbf{24} (1972), 119--148.
6944: %
6945: \bibitem{vA} W.\ Van Assche,
6946: \emph{Rakhmanov's theorem for orthogonal matrix polynomials on the
6947: unit circle},
6948: J.\ Approx.\ Theory \textbf{146} (2007), 227--242.
6949: %
6950: \bibitem{V35} S.~Verblunsky,
6951: \emph{On positive harmonic functions: A contribution to the algebra
6952: of Fourier series},
6953: Proc.\ London Math.\ Soc. (2) \textbf{38} (1935), 125--157.
6954: %
6955: \bibitem{Weyl} H.~Weyl,
6956: \emph{\"Uber gew\"ohnliche Differentialgleichungen mit Singularit\"aten
6957: und die zugeh\"origen Entwicklungen willk\"urlicher Funktionen},
6958: Math.\ Ann. {\bf 68} (1910), 220--269.
6959: %
6960: \bibitem{Whit} P.~Whittle,
6961: \emph{On the fitting of multivariate autoregressions and the
6962: approximate canonical factorization of a spectral density matrix},
6963: Biometrika \textbf{50} (1963), 129--134.
6964: %
6965: \bibitem{Wid} H.~Widom,
6966: \emph{Polynomials associated with measures in the complex plane},
6967: J.\ Math.\ Mech. {\bf 16} (1967), 997--1013.
6968: %
6969: \bibitem{Wie} N.~Wiener,
6970: \textit{Extrapolation, Interpolation, and Smoothing of Stationary
6971: Time Series. With Engineering Applications},
6972: The Technology Press of the Massachusetts Institute of Technology,
6973: Cambridge, Mass. (1949).
6974: %
6975: \bibitem{Wong} M.-W.~L.~Wong,
6976: \emph{First and second kind paraorthogonal polynomials and their zeros},
6977: J.\ Approx.\ Theory \textbf{146} (2007), 282--293.
6978: %
6979: \bibitem{Yak} H.~O.~Yakhlef,
6980: \emph{Relative asymptotics for orthogonal matrix polynomials with
6981: unbounded recurrence coefficients},
6982: Integral Transforms Spec.\ Funct. \textbf{18} (2007), 39--57.
6983: %
6984: \bibitem{YM} H.~O.~Yakhlef and F.~Marcell\'an,
6985: \emph{Orthogonal matrix polynomials, connection between
6986: recurrences on the unit circle and on a finite interval},
6987: Approximation, Optimization and Mathematical Economics
6988: (Pointe-\`a-Pitre, 1999), pp.~369--382, Physica, Heidelberg (2001).
6989: %
6990: \bibitem{YM02} H.~O.~Yakhlef and F.~Marcell\'an,
6991: \emph{Relative asymptotics for orthogonal matrix polynomials with
6992: respect to a perturbed matrix measure on the unit circle},
6993: Approx.\ Theory Appl.\ (N.S.) \textbf{18} (2002), 1--19.
6994: %
6995: \bibitem{YMP01} H.~O.~Yakhlef, F.~Marcell\'an, and M.~A.~Pi\~{n}ar,
6996: \emph{Perturbations in the Nevai matrix class of orthogonal matrix
6997: polynomials},
6998: Linear Algebra Appl. \textbf{336} (2001), 231--254.
6999: %
7000: \bibitem{YMP02} H.~O.~Yakhlef, F.~Marcell\'an, and M.~A.~Pi\~{n}ar,
7001: \emph{Relative asymptotics for orthogonal matrix polynomials with
7002: convergent recurrence coefficients},
7003: J.\ Approx.\ Theory \textbf{111} (2001), 1--30.
7004: %
7005: \bibitem{YK} D.~C.~Youla and N.~N.~Kazanjian,
7006: \emph{Bauer-type factorization of positive matrices and the theory of
7007: matrix polynomials orthogonal on the unit circle},
7008: IEEE Trans.\ Circuits and Systems {\bf CAS-25} (1978), 57--69.
7009: %
7010: \bibitem{Zyg} M.~Zygmunt,
7011: \emph{Matrix Chebyshev polynomials and continued fractions},
7012: Linear Algebra Appl. \textbf{340} (2002), 155--168.
7013: 
7014: 
7015: \end{thebibliography}
7016: 
7017: \bigskip
7018: \noindent David Damanik \\
7019: \noindent Department of Mathematics \\
7020: \noindent Rice University \\
7021: \noindent Houston, TX 77005, USA \\
7022: \noindent damanik@rice.edu
7023: 
7024: \bigskip
7025: \noindent Alexander Pushnitski \\
7026: \noindent Department of Mathematics \\
7027: \noindent King's College London \\
7028: \noindent Strand, London WC2R 2LS, UK \\
7029: \noindent alexander.pushnitski@kcl.ac.uk
7030: 
7031: \bigskip
7032: \noindent Barry Simon \\
7033: \noindent Mathematics 253--37 \\
7034: \noindent California Institute of Technology \\
7035: \noindent Pasadena, CA 91125, USA \\
7036: \noindent bsimon@caltech.edu
7037: 
7038: 
7039: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7040: \endddoc
7041: %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%
7042: 
7043: %\end{document}
7044: